Sei sulla pagina 1di 49

Chem. Rev.

1907, 8 7 , 433-481

433

Photochemical Reactions of Organic Crystals


V. RAMAMURTHY and K. VENKATESAN
Department of Organb Chemistry, Indian Institute of Science, Bangalore-560 0 12, India
Received April 3, 1986 (Revised Manuscript Received October 13, 1986)

C. Consequences of Local Stress on

Contents
I. Introduction

433

I I. Topochemical Principles: Correlation of


Structure with Reactivity in Bimolecular

Reactions
A. Cycloaddition Reactions
1. Distance Criteria
2. Parallelism of Double Bonds
3. Minimum Translational Movement in the
Crystal Lattice
4. Single Crystal to Single Crystal
Photodimerization
5. Solid-state Asymmetric Synthesis
6. Competlng Dimerization Reactions
7. Miscellaneous Dimerization Reactions
111. Topochemical Postulate and Unimolecular
Transformations
A. Intramolecular Hydrogen Abstraction
Reactions

434

436
438
440
44 1
44 1
443
445
446
447

45 1

5.

P,y-Unsaturated Ketones
Aryl Ketones
Nitroaromatics

454

6.

Imines

456

B. Fragmentatlon Reactions
C. Electrocyclization
D. Photochemical Oxygen-Transfer Reactions
E. Miscellaneous Reactions
F. Asymmetric Synthesls
IV. Structure-Reactivity Correlations in Gas-Solid
Photoreactions
V. Crystal Engineering
A. Chloro Substitution
B. Methyl Substitution and the Question of
Isomorphism with Chloro Derivatives
C. Acetoxy Substitution
D. Methylenedioxy Substitution
E. Complexation with Lewis Acids
F. Solid Solution or Mixed Crystal Formation
G. Unimolecular Reactions
VI. Subtler Aspects of Photoreactivity in Solids
A. Role of Neighbors in Solid-State Reactions:
Concept of Reaction Cavity
B. Photoinduced Lattice Instability: Concept
of Dynamical Preformation

475
477

I . Intfoduct/on

447

4.

475

435

1. Tetrahydronaphthoquinones
2. Tetrahydronaphthoquinols
3.

Solld-State Reactions
D. Role of Defects
VII. Conclusion

449
452

457
458
459
460
462
462
464
465
466
467
467
468
468
469
470
470
473

0009-2665/87/0787-0433$06.50/0

Organic photoreactions in the crystalline state have


been studied widely and date back to the end of the last
century.lg Early organic chemists frequently worked
with solid materials because of the unavailability or
expense of suitable solvents. But growth came to a halt
because not much was known at that time of the nature
and structure of crystals. However, as the years passed,
synthetic and mechanistic organic chemists have concentrated on reactions in fluid media and solid-state
reactivity did not enjoy in these years the popularity
of solution reaction studies. The main experimental
obstacle was probably the difficulty of identifying reactive crystals. In order b control and exploit organic
reactions in crystals, chemists must develop the same
sort of intution about these processes that they have
developed for fluid-phase reactions.
The science of solid-state organic chemistry, and
particularly the area of lattice control over reaction
pathways, now seems to be entering a period of flowering and growth. There is no doubt that with deeper
understanding of packing effects and of bpochemistry,
solid-state organic reactions could be planned and exploited in organic chemistry. The combination of organic solid-state chemistry and X-ray crystallography
has proved to be invaluable for structure-reactivity
correlation studies. Techniques such as X-ray crystallography, high-resolution electron microscopy, and
solid-state magic angle spinning NMR spectroscopy
have opened up entirely new dimensions in organic
solid-state chemistry. Although considerable progress
has been made in this multi-disciplinary field in the last
two decades, much chemistry still remains to be done.
It seems appropriate at this stage of development in this
field to review in depth the existing literature as it
stands in 1986. This review is primarily aimed at the
practicing organic chemist.
The article is primarily concerned with organic
photoreactions in the crystalline state. Photopolymerization in the crystalline state is not covered.
It is important to note that tremendous progress has
been made in this area during the last
Studies on inclusion complexes and on surfaces have
not been included although interesting results have
been r e p ~ r t e dfurther,
;~
photophysical studies of organic
crystals have also been excluded. Photodimerization
of cinnamic acid and its derivatives in polymer matrix
0 1987 American Chemical Society

434 Ch3mlcai Reviews. 1987. Vd. 87,

Rammuthy and Venkatesan

No. 2

reactions in crystals and stimulate further explorations


in this area. A number of other reviews relevant to this
topic have been published earlier.14

I I . Topochemkal Principles: Correlation of


Structure with Reactivity in Bimolecular
Reactions

V. Ramamurthy obtained his M.Sc. degree (1968) from I.I.T..


Madras. and the h . D . &gee (1974) from the Universw of Hawaii
(Prof. R. S. H. Uu). Following posMoctwal studies at the Unhrenny
of Western Ontarb (with P. de Mayo; 1974-75) and Columbia
University (with N. J. Turro; 1975-78). he joined the Department
of Organic Chemisby. Indian Institute of Science. where he is
currently Associate Professor. During 1978-1985 he has spent
brief periods as a Visiting Scientist at the University of British
Columbia. the University of Notre Dame. Columbia University. the
University of Western Ontario. and the University of Hawaii. He
is a Fellow of the Indian Academy of Sciences and a Research
Fellow of the Indian National Science Academy (1985-1987). His

research interests lie principally in organic photochemistry.

K. Venkatesan received his w


e
e in physics in 1953 at Annamalai
ulhnnii and a ph.D under the wpmislon of Rot. S. Ramaseshan
at the Indian Instiie of Scienat. He cad& ouf postdoncral work
wHh Prof.J. D. Dunitz at tha Swiss Federal Instiiute of Technolosy
and Prof.Dcfothy Crowfoot Hodgkin at Oxford University. He was
on the Facuity of tha physics Department of Madras Univmity from
1963 to 1971. I n 1971 h e joined the Facuity of the Department
of Organic Chemistry at the Indian lnstiiute of Science. becoming
a Professor in 1981. His main research interests are solid-state

chemistly and chemical crystallography.

is not included. Although these are the first known and


first used negative photoresist, as the medium of reaction is not crystalline state their inclusion appeared
inappropriate. This review will attempt to illustrate the
main features of organic solid-state photoreactions with
a range of appropriate examples. It is hoped that this
treatment will demonstrate the utility of conducting

Schmidt and co-workers at the Weizman Institute,


studied systematically (in the early 1960s) the factors
that govem the course of organic solid state, especially
photoinduced, reactions.17 As a result of their extensive studies on the photodimerization of cinnamic
acids, they confirmed the topochemical postulate first
enunciated hy Kohlschutter in 1918,*which states that
reactions in crystals proceed with a minimum of atomic
and molecular movement. According to Kohlschutter,
a topochemical reaction is one in which both the nature
and properties of the products of the reaction are governed by the fact that it takes place under the constraining influence of the three-dimensional periodic
environment. As a result of the work of Schmidt and
co-workers some important principles have been established and these are discussed in this section. The
important points to emerge and those that form the
basis of topochemical control of both bimolecular and
unimolecular transformations are the following:
(i) The intrinsic reactivity of a molecule is less important than the nature of the packing of neighboring
molecules around the reactant.
(ii) The separation distance, mutual orientation, and
space symmetry of reactive functional groups are crucial.
(iii) In crystalline solids there are very few (usually
just one) conformations taken up by molecules which,
in the dispersed state, are very flexible.
(iv) Molecular crystals (into which category the vast
majority of organic solids fall) display a rich diversity
of polymorphic forms, in each of which a particular
conformer or particular symmetry and separation of
functional groups prevails.
Following the pioneering investigations of Schmidt
and his co-workers several groups have attempted to
understand organic solid-state transformations on the
basis of crystal structure. In this section, we highlight
the role of solid-state structure in controlling solid-state
reactions. Both unimolecular and bimolecular organic
photoreactions are covered. Where crystal structures
are available, an attempt is made to relate the structure
with the reactivity.
The reactions of trans-cinnamic acids in the crystalline state are well known examples of [2 + 21 photodimerization and the classic studies by Schmidt and
his co-workers have demonstrated that such reactions
are strictly controlled hy the packing arrangement af
molecules in the crystals. Since the original contributions by Schmidt and his group, several examples
of photodimerization in the solid state have been reported but these have been isolated cases rather than
forming a systematic investigation with the emphasis
on structure reactivity correlations. In this context
recent investigations on coumarinsl*u and benzylidenqclopentanonessm are noteworthy. A large number
of closely related molecules in these two series have
been subjected to both photochemical and crystallo-

Chemical Revlews, 1987, Vol. 87, No. 2 435

Photochemical Reactions of Organic Crystals

SCHEME 1
TOPOCHEMICAL

PWOTODIMERIZATION

Double bond s e p a r a t i o n

: 3.6 -4.l.A

%h
coon
d-TRUXILLIC ACID

ph

h3

coon Solid

Oh

Ph

p-

form (3.9-4.1.A;

Translation 1

9 C O O I - I

coon

solution

/J-TRLJXINIC
r-form

(4.7- 5.1.A;

Translation )

ACID

NO REACTION

Reaction i n the solid s t a t e o c c u r s w i t h o m i n i m u m amount of o t o m i c o r molecular movement


G.M.J. S c h m i d t

graphic studies. While studies on cinnamic acids resulted in very important correlations between molecular
alignment in the reactant crystal and steric configuration of the product, analyses of the behavior of benzylidenecyclopentanones and coumarins in the solid state
have provided an opportunity to reexamine the subtler
aspects of the topochemical postulates. The following
section is devoted to a brief discussion of the important
factors which are crucial for topochemical dimerization
in the solid state. Other known examples of dimerization of organic molecules (often unrelated) in the solid
state are summarized in the form of schemes.

(1964 -1971

VIEW "X"

Figure 1. Stereodrawing of the a-packing of trans-cinnamic acid.

A. Cycloaddition Reactions
VIEW "Y"

The reactions of cinnamic acids are examples of [2


+ 21 photodimerization which have been investigated
extensively. Some of these acids, on photolysis of the
crystal, react to give dimeric products (Scheme 1) while
in solution trans-cis isomerization occurs but there is
no dimerization. The acids are observed to crystallize
in three polymorphic forms, namely, a,p, and y, and
show photochemical behavior which is determined by
this structure type. In all three modifications, cinnamic
acid molecules pack in one dimensional stacks, adjacent
stacks being paired by hydrogen bonding across centers
of symmetry.'"l7 Within the stacks the molecules lie
parallel with the normal distance between molecular
planes being of the order of = 3.5 A. The three
structural types differ in the angle that the stack axis
makes with the normals to the molecular planes. This
is equivalent to a difference in the distance between
equivalent points on the molecules, which is the crystallographic repeat distance, "d". In the &type structure the molecules are separated by a short repeat
distance of 3.8-4.2 A, thus neighboring molecules up the
stack are translationally equivalent and show considerable face to face overlap. All cinnamic acids which
crystallize in this structure react photochemically to
give products of the same stereochemistry (mirror
symmetric dimers). In the y-type structure adjacent
molecules are offset so that the reactive double bonds
do not overlap, and furthermore the distance between

Figure 2. Stereodrawing of the P-packing of p-chloro-transcinnamic acid.

them is large (4.8-5.2 A). Crystals of this type are


photostable. In the a-type the double bond of a molecule in one stack overlaps with that of a centrosymmetrically related molecule in an adjacent stack. The
distance between the equivalent double bonds is greater
than 5.5 A, but that between the overlapping double
bonds is ~ 4 . A.
2 This type of crystal upon irradiation
produces centrosymmetric dimers. The examples of a,
p, and y packing are shown in Figures 1-3.
Based on extensive crystallographic and photochemical studies on cinnamic acids Schmidt deduced the
following conclusions: (a) The nature of the crystal

436 Chemical Reviews, 1987, Vol. 87,

No. 2

Ramamurthy and Venkatesan

6-

SCHEME 3

Figure 3. ?-Packing of coumarin.

*oy+No2
R

SCHEME 2
R
MeOOC
COOMe

h3
solid

*
R
R = COOMe
CN

CN

h3
solid

-CN

1. Distance Criteria

*
CN
COOMe

MeOOC
-COO

h3
H

solid

MeOOC

QCOOH
COOH

CONH2

H2NOC
=CONH,

h3
solid

Itr

HzNOC
~

rationalized on the basis of Schmidt's criteria for dimerization.


However very recent studies on the photodimerization of olefinic crystals have brought out several examples which deviate significantly from the
well-accepted topochemical principles. Below we shall
dwell upon such exceptional cases in the light of the
original topochemical principles for dimerizations.

CONH2

CONH2

structure determines whether or not reaction occurs and


the molecular structures of the products, if any. (b) The
reaction involves combination between nearest neighbor
molecules in a stack, and occurs with a minimum of
atomic and molecular movement. Schmidt has drawn
attention to the fact that not only must the double
bonds of the reacting monomers of cinnamic acid be
within frl 4.2 A, but they must also be aligned parallel
for cycloaddition to occur. A reaction which behaves
in this way is said to be "topochemically controlled".
Schmidt has drawn the geometrical criteria for dimerization only with the view of inferring how precisely the
7r electron system of the reacting double bonds must be
aligned in the crystal lattice for reaction to occur.
These topochemical postulates are landmarks in organic solid state photochemistry and are used as rules,
as they are able to provide an understanding of a large
number of [2 + 21 photodimerization reactions of widely
varying structures. For example, dimerization of fumaryl derivatives3' (Scheme 2), heterocyclic analogues
of trans-cinnamic acid (Scheme 3),32butadiene derivatives (Scheme 4),33,34
c o u m a r i n ~ and
, ~ ~benzylide~~
necyclopentanones2"30 (Scheme 5 and 6) have all been

As presented above, for dimerization to occur the


distance between the potentially reactive double bonds
should be less than 4.2 A. With the exception of methyl
p-iodocinnamate, all the cinnamic acid derivatives
which have adjacent double bonds separated by a distance of more than 4.2 A in the crystalline phase are
ph0tostab1e.l"~~In the case of methyl p-iodocinnamate,
the molecules are arranged in a &type packing with an
interdouble bond distance of 4.3 A and yet react to yield
the expected photodimer.15-17
However, the upper limit of the critical distance for
photodimerization in the solid state is not absolutely
established, as the limit was set in the absence of further experimental data in the range 4.2-4.7 A, above
which photodimerization does not occur. The results
with 7-chlorocoumarin are of interest in this context.23
Irradiation of crystalline 7-chlorocoumarin yielded a
single dimer (syn head-head), without induction period,
in -70% of yield (Scheme 5 ) . The packing arrangement reveals that the two potentially reactive 7chlorocoumarin molecules are separated by 4.45 A, this
being the repeat along the a-axis. Since the only dimer
obtained corresponds to syn head-head, it is clear that
the reaction is between the pairs translated alon the
a-axis. It is noteworthy that the distance of 4.45 lies
beyond the so far accepted limit of 3.5-4.2 A for photodimerization in the solid state. Very recently photodimerization of retinoids in the solid state has been
reported.35 Etretinate (1) dimerizes in the solid state
to yield two dimers (Scheme 7). The center-to-center
distance for the two sets of dimerizable bonds are 3.8
and 4.4 A, the latter being outside the presently accepted limit. The most unusual case reported so far is
p-formylcinnamic acid.36 This crystal, possessing a
b-axis of 4.825 A, dimerizes in the solid state to yield
a mirror symmetric dimer. It may however be noted
that the plane to plane perpendicular distance between
reactive molecules is fairly short (3.88 A).37 Thus the

Chemical Reviews, 1987, Vol. 87, No. 2 437

Photochemical Reactlons of Organic Clystals

NC

-CN

h3

NC

CN

CN

Ph
Ph
-CONHp

Ph

SCHEME 6

SCHEME 5

CqlJQfX"
0
(anti head- head I

(svn head-head)

Coumarin Irradiated

- MethoZy
7 MoihoUy
8 - MethOXY
6 ACetOXy
7 - Acetouy
4 Methyl ?-ACotOXY
4 Chloro
6 Chloro
? Chloro
4 Methyl 6 - Chlaro
4 Methyl 7 - Chloro
7 Methyl
6

(anti head-tail)

0
X

X = H
: Br

( s y n head-tall)

h3

Dimer

Nature of Dimor

syn H H (60%)
HT (90~01
HT ( 5 0 % )

syn
ant1
syn
syn

H H (70%)
HH (90%)
syn HH (eo%)
anti HH and syn HT (25'10)
syn H H (100%)
syn HH ( 7 0 % )
syn HH (50%)
syn HH ( 8 0 % )
syn H H (65%)

BP-CI B C P
Bp-Me BCP

Y =HI
Y = H,

X :p-CI
X = p-Me

p-Me B B C P

Y = p-CI, X
Y =p-Met X

p-CI 8 p - B r B C P
p-MeBp-Br B C P

Y zp-13, X = p-Br
Y:p-Me, x = p-Br

p-CI B B C P

H
:

No
NO

Yes} Centrosymmetric
dimer
Yes
NO
YOS

438 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

SCHEME 7

-1
9 0O/O

COOEt
Et OOC
10O/O

SCHEME 8

COOCH3

Hon/Et A
Et OOC

iO*

X = COOCH3

-2

-3
0
-N.FNg0

H0b

6
-

-7

above examples point out the need for a closer examination and modification of the distance criteria for
photodimerization.
The opposite situation, namely, absence of reaction
in spite of favorable distance, has also come to light and
consideration of this allows us to know more about the
subtler aspects of photoreactivity in crystals. One of
the polymorphs of distyryl pyrazine where the potentially reactive double bonds are separated by 4.19 A is
photostable.% The photostability of this compound has
been ascribed to the layered structure which suppresses
the molecular deformation necessary for the cycloaddition reaction. Another example where the molecular packing satisfies the topochemical criteria but yet
is photostable is 239 (see Scheme 8 for nonreactive
olefins). The potentially reactive double bonds are
parallel with a center-to-center distance of 3.79 A.
Nevertheless., 2 is photochemically inert when irradiated
in the solid state. The probable reason for the lack of
solid-state reactivity of this enone is the steric compression experienced by the reacting molecules at the
initial stages of photocycloaddition (discussed in section
VI). In the crystal of methyl 4-hydroxy-3-nitrocinnamate (3), the nei hboring molecules are related by
a translation of 3.78 . But it has been observed that
this compound is photostable in the solid state.Q41 In
the crystal structure the molecules are linked by hydrogen bonds to form a sheetlike structure close to the
(102) plane. It is likely that the extensive intermolecular hydrogen bond network and C-He-0 type interactions involving the ethylenic carbon atom do not

permit the easy spatial movement of the atoms of the


double bond in the lattice for the reaction to proceed.
It has been reported that in the photodimerization of
diethyl succinylsuccinate (4) the conversion of monomer
to dimer is only 3%.42943
However, the crystal structure
of this compound reveals that the double bonds are
separated by 3.58 A and are conducive for photodimerization. It has been observed in our laboratory
that benzylidene-dl-pipertone(5) is photostable in spite
of the fact that there are two pairs of centrosymmetrically related double bonds which are parallel and at
a distance of 3.92 and 3.98 A, respectively.4 Crystalline
(+)-2,5-dibenzylidene-3-methylcyclopentanone
(6) and
2-benzylidenecyclopentanone (7) are photostable while
closely related systems possessing similar packing arrangements undergo dimerization readily in the solid
state (Scheme 9).&q6 The distance between the centers
of the olefinic bonds of the inversion related pairs in
the former and in the latter are 3.87 and 4.14 A, respectively. The photostability is attributed to the reduced overlap between potentially reactive orbitals. As
we shall see in a later section the absence of photoreactivity in many of these cases can be understood by
performing lattice energy calculations.
2. Parallelism of Double Bonds

According to Schmidt and co-workers, parallel


alignment of double bonds is as important as the distance between them. For example, methyl m-bromocinnamate does not yield any photodimer even on long

Photochemical Reactions of Organic Crystals


METHYL

n-mono

CINNFIMFITE --INITIFIL

Chemical Reviews, 1987, Vol. 87, No. 2 439


ORIENTFITION

SCHEME 10

solid

0
CH3 Ph Ph

02%

Figure 4. Stereodrawing of the packing of methyl m-bromocinnamate.


SCHEME 9

0
b

solid
h3

solid
h3 * P

s=T

- Syn - polymer

No r e a c t i o n

10

UV exposures. The distance between the centers of the


adjacent double bonds is 3.93 A, but the double bonds
are not parallel. The double bonds make an angle of
28 when projected down the line joining the centers
of the bonds (Figure 4). In the crystal of 1,l-trimethylenebis(thymine) (8) the thymine rings are arranged such that both intermolecular and intramolecular photoreactions could occur.47 Irradiation of this
compound yields, however, a polymeric product which
would arise from intermolecular reaction. This is because the intermolecular double bonds are nearly parallel to each other, while the intramolecular double
bonds are inclined to each other at an angle of
6.
Yet another convincing example is provided by [2,2](2,5)benzoquinophane (9).48 Here both inter- and intramolecular cycloadditions are favored according to the
distance criteria (-3.80 and -3.0 A, respectively).
However, photolysis yields only intramolecular addition.
This is attributed to the fact that the intermolecular
adjacent double bonds are inclined to each other at an
angle of 53.3.
On the other hand, a few cases have also been reported where exact parallelism between reactant double
bonds has not been adhered to and yet photodimerization occurs. For example, in the crystals of
7-methoxycoumarin, the reactive double bonds are rotated by about 65 with respect to each other, the
center-to-center distance between the double bonds
being 3.83 A.24 In spite of this unfavorahle arrangement, photodimerization occurs giving syn head-tail

dimer as the only product in quantitative yield (Scheme


5). 1,4-Dicinnamoylbenzene (11) crystals undergo
photodimerization via a double [2 + 21 cycloaddition
in a topochemical manner to the corresponding cyclic
product (Scheme
The nearest monomer units
are arranged skew to each other (6) and the distances
between intermolecular double bonds are 3.97 and 4.09
A for one cyclobutane ring and 3.90 and 3.96 A for the
other. 2,5-Dibenzylidenecyclopentanone10 is analogous
in its behavior and packing to 7-methoxy~oumarin.~~
When 10 is irradiated by UV light in the crystalline
state the principal product is formed by a [2 + 21 dimerization (Scheme 9). The cyclopentanone 10 molecules are arranged such that the mean distance separating the potentially reactive centers is -3.7 A, the
angle between the two bonds being 56. Although this
is not the geometry considered conducive for a topochemical reaction, dimerization does indeed take place
in the solid state.
A few examples of other closely related reactions
wherein the reactive olefinic ?r systems are skewed with
respect to each other are also reported. In the [4 + 41
photocycloaddition of l-methy1-5,6-diphenylpyrazin-2one (121, it has been reported that even though the
double bonds are twisted by an angle of 24 (the angles
between C2=Cs and N2=C4), dimerization does occur
(Scheme
In this case, the distances between the
double bonds are 3.46 and 3.64 A, which are well within
the proposed limit. Another case in which the reaction
takes place between nonparallel double bonds is the
photopolymerization of the diacetylene, 2,4-hexadiyne
126-diylbis(m-tolylmethane)
(13) where the closest distance of approach of triple bonds is 3.61 A.51 In this
crystal although the reactive triple bonds are crossed
at an angle of 72 polymerization does occur (Scheme
10). 1,l-Trimethylenebis(thymine)undergoes photopolymerization via [2 + 21 cycloaddition in a topo-

Ramamurthy and Venkatesan

440 Chemical Reviews, 1987, Vol. 87, No. 2

TABLE 1. Relative Orientations of Reactive Double Bonds i n Coumarins"


center-to-center
distance between
the reactive
coumarin
double bonds, A
81, deg
7-chlorocoumarin
4.45
0
pair I (translation)
pair I1 (centrosymm)
4.12
0
4-methyl-7-chlorocoumarin
4.08
0
7-acetoxycoumarin
0
3.83
8-methoxycoumarin
pair I
4.07
0
pair I1
0
3.86
7-methoxycoumarin
3.83
65
ideal values
0
4.2

deg

e3, deg

displacement of
double bonds upon
projection, A

131.4
127.9
121.4
106.4

85.3
107.0
88.53
125.45

0.287
0.936
0.011
1.329

122.4
117.4

63.77
112.88

1.565
1.333

82,

90

90

0.0

"For a definition of geometrical rmameters see Figure 5.

chemical manner as described earlier.47 Ever1 in this


case the double bonds are twisted with respect to each
other by 5'.
The anomalous examples presented in this and earlier
sections raise certain important questions concerning
the correlation of reactivity with structure. It is important to stress that Schmidt's original criteria have
explained a large number of topochemical dimerization
reactions. Exceptions observed in recent studies should
not be construed as serious violations of original concepts but be integrated into the original basic ideas by
widening apparent limitations and scope.

3. Minimum Translational Movement in the Crystal


Lattice

The topochemical principle states that reaction in the


solid state is preferred and occurs with a minimum
amount of atomic or molecular movement. This implies that a certain amount of motion of various atoms
in the crystal lattice is tolerable. Based on this, one
could assume that for the formation of a cyclobutane
ring with C-C lengths of 1.56 A the double bonds can
undergo a total displacement of about 2.64 %I toward
each other from the original maximum distance of 4.2
A. Even under ideal conditions, movement of double
bonds toward each other is essential for dimerization
to take place. The criterion of less than 4.2 A separation
implicitly assumes that such a motion can be accommodated in the crystal. It would be expected that in
some cases molecular motion such as (i) rotation of
double bonds with respect to each other (to bring about
parallelism from a nonparallel arrangement), (ii) a rotation about its own C=C axis (to achieve a maximum
overlap of the K orbitals), (iii) translation of double
bonds in the plane of the molecule, and (iv) movement
along the C-C double bond axis may become necessary
before dimerization can take place. We discuss below,
based on the photodimerization of coumarins, that the
types of motions mentioned above may indeed occur
in the crystalline state.
Geometrical parameters that are useful in addition
to center-center distance are el, e,, e,, and the displacement of double bonds with respect to each other
(Figure 5).19 B1 corresponds to the rotational angle of
one double bond with respect to the other, 13, corresponds to the obtuse angle of the parallelogram formed
by double bond carbons C3, C4,C3', and Ci,whereas O3

Figure 5. Geometrical parameters used in the relative representation of reactant double bonds.

measures the angle between the least square plane


through the reactive bonds C3,C4,C3/, and Cq/ and that
passing through C',, C3, C4,and C'lo. The basic aim in
performing these rotations is to bring the K orbitals of
the reacting partners to overlap, the ideal values for el,
d2, and e3 for the best overlap being 0, 90, and 90,
respectively. While 8, reflects the displacement along
the double bond axis, O3 is a memure of ita displacement
in the molecular plane. Perusal of Table 1reveals that
in all the four reactive coumarins the reactive double
bonds are not ideally placed. Although they are coplanar and parallel, the two double bonds are displaced
with respect to each other both in the molecular plane
as well as along the double bond axis. In all the four
cases the configuration of the dimers obtained corresponds to the one that is expected based on molecular
packing in the crystal. This suggests that motions of
the molecules in the molecular plane and along the
double bond axis are required and do indeed occur in
the crystal lattice. Thus a certain amount of flexibility
in the motions of the molecules in the crystal is to be
expected. In this context a few new concepts have ,
emerged in recent times and these will be discussed in
section VI.

Photochemical Reactions of Organic Crystals

Chemical Reviews, 1987, Voi. 87, No. 2 441

TABLE 2. Crystallographic Data Associated with the Single Crystal


Single Crystal Dimerization of
2-Benzyl-5-benzylidenecyclopentanone(BBCP) and its p -Bromo Derivative (BpBr-BCP)
compda
cell parameters
monomer
dimer
70 change

BBCP

a, A
b, A
c, A

z
BpBr-BCp

space group
unit cell volume, A3
a, A
b, 8,
c, A

space group
unit cell volume, A3

31.30
10.78
8.69
8
Pbca
2932
34.25
10.88
8.43
8
Pbca
3141

31.32
10.81
8.63
4
Pbca
2922
32.96
10.27
8.98
4
Pbca
3040

0.06
0.28
-0.69

-3.77
-5.61
-6.52

For structure of compounds see Scheme 6.


TABLE 3. The Most Common Space Groups of Molecular
Crystals Based upon a Survey of Some 5000 Crystal
Structure Determinations
mace erouo
number
oercentane
E l I C

%421
Pi
E 1

c2/c
Pbca

Figure 6. Composite diagram comparing the packing of the


molecular units within the monomer and dimer crystal structures
of BBCP.

4. Single Crystal to Single Crystal Photodimerization

There is cogent evidence to suggest that for certain


photoinduced and thermally induced reactions the nature and stereochemistry of the product is precisely
determined by the crystal packing within the perfect
monomer lattice. However, the detailed mechanism by
which organic reactions proceed within the solid state
is not completely understood. Although in early 60s
Schmidt alluded to the mechanistic details of photodimerization and suggested the possibility of occurrence
of single crystal to single crystal dimerization, such
examples were discovered only recently. Single crystal
to single crystal polymerization of diacetylene constitutes the first example of such reactions.52 Photodimerizations of 2-benzyl-5-benzylidenecyclopentanone
(BBCP) and its p-bromo derivative (14 and 15) have
also been established to be single crystal to single crystal
transformations.26vmrBcrystallographic details provided
in Table 2 illustrate the similarity in cell dimensions
between monomers and dimers. Further, packing arrangements shown for BBCP and its dimer reveal that
the dimerization process requires very little motion of
the atoms (Figure 6). Most interestingly, the specific
atoms involved in the reaction have been directly
monitored during the dimerization through X-ray diffractometer measurements.
5. Solid-state Asymmetric Synthesis

The achievement of an asymmetric synthesis starting


from an achiral reagent and in the absence of any external chiral agent has long been an intriguing challenge

1897
839
449
418
310
247

37.9
16.8
9.0

8.4
6.2
4.1

to chemists. The results obtained by Schmidt and coworkers on cyclodimerization in the crystalline state
with well defined stereochemistry led them to the next
logical step of achieving asymmetric synthesis in the
solid state. Before describing their results it would be
appropriate to mention a few facts relating to the symmetry of crystals. There are 230 genera of space groups
which can be divided into two categories: (a) the chiral
space groups, 65 in number, have only symmetry elements of the first kind, Le., translations, rotations, and
combinations of these; (b) the nonchiral space groups,
of which there are 165, may contain symmetry elements
such as a mirror plane or glide plane or center of inversion. Thus the unit cell of a compound belonging
to an achiral space group will contain both the object
and its mirror image. It is obvious that any attempt
at achieving asymmetric synthesis via photochemical
reactions should begin with a compound crystallizing
in any one of the 65 chiral space groups. As to be
discussed in section V, crystal engineering is not so
advanced that any desired crystal environment can be
prepared to order. In this connection the following
observations are worthy of mention. In a survey of
some 5000 X-ray structure determinations of homomolecular crystals reported (with data published in
1924-1975) it has been observed that organic molecules
tend to crystallize in the systems of low symmetry
namely, monoclinic and orthorhombicsystems.53 Of the
219 distinct space groups (11enantiomorphous groups
excluded from the total of 230) the most commonly
occurring space groups (Table 3) are E 1 / c , P212121, PI,
E,,
C2/c and Pbca, the chiral ones being P2,2,2, and
El.
Yet another crystallographic observation of relevance for asymmetric synthesis in crystals is the phenomenon of achiral molecules crystallizing in chiral
space groups. In these instances (e.g. benzophenone,
y-modification of glycine and binaphthyl) the chiral
environment of the crystal forces the molecule to acquire a chiral conformation.

442

Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

Ph
P

h \ w Ar

16

Ph

Ar

solid

le

T%Ar

Ar

p%Ar

Ar

L7
P

Ar
7H3

Ar

,OCHCHZCH~
C*O ,OCH~

Ar

c H3

OCHCHzCH3

Figure 7. Asymmetric synthesis using mixed crystals of 16 and


17.
cO,
CH~O'

>c

CH3CH2CHO

CH3

CH CH CHO
21
CH3

Figure 9. Schematic representation of the asymmetric dimerization of 21.


SCHEME 12
X

Figure 8. Stereodrawing of the intermolecular overlap in mixed


crystals of 16 and 17.
SCHEME 11
X

-'

zI

Y
I

solid
h3

-@+@
I

On the basis of the vast amount of knowledge gained


regarding photodimerization reactions in the solid state,
the Weizmann group extended their studies to design
and perform solid state asymmetric syntheses.%* Two
approaches were mainly used in this venture. The enantiomeric yields achieved in these reactions varied
from a few percent to quantitative. In the first approach the basic idea is to grow crystals containing two
components (Scheme 11). It was known from their
earlier studies that phenyl-substituted olefins (stilbenes
and 1,4-diarylbutadienes) tend to adopt chiral structures. Further there were clear indications that dichloro
derivatives bring about a 4-A packing arrangement. As
for the choice of the two components, thiophene and
phenyl derivatives were favored since these molecules
have been found to form mixed crystals (substitution
solid solution). Thus when mixed crystals of 16 and 17
(Figure 7) are irradiated, enantiomeric heterodimen (18
and 19), in addition to the meso homodimers, are
formed. In actual practice mixed crystals containing
85% of 16 and 15% of 17 were used yielding heterodimers in high percentage. Figure 8 represents the
arrangement of host 16 and guest 17 molecules in the
crystal lattice along the 4 A axis. The absolute configuration of the mixed crystal has been established
using X-ray crystallography.
The strategy adopted in the second approach de-

manded that molecules having two nonidentical reactive


sites pack in a chiral crystal in such a way that the
nonequivalent double bonds overlap. Such an arrangement would be expected to generate only one of
the two possible enantiomers (Scheme 12). The actual
system chosen for implementing this proposal was
benzene-l,4-diacrylates.The chiral sec-butyl group was
used to induce the compound to crystallize in a chiral
space group. When the ethyl ester derivative (S)-(+)-20
was irradiated at 5 "C, chiral dimer, trimer, and higher
polycyclobutane oligomers were obtained in high yields
approaching 100%. Irradiation of the enantiomeric
(R)-(-)-20isomer yielded products having optical rotations of the same magnitude but of opposite sign.
That the chiral induction is entirely dictated by the
chirality of the crystal environment and owes not to the
presence of the chiral sec-butyl group was also established. The enantiomeric methyl ester monomer (S)(+)-21is dimorphic. Form one is light stable whereas
form two is photoactive yielding dimers 22 and 23
(Figure 9). Although the monomer crystal is chiral, the
irradiation product, after removal of the sec-butyl
group, turns out to be racemic. This is explained in
terms of the crystal packing of monomer in which the
monomers are related by a pseudocenter of inversion
(pseudo P2,la). There is also an interesting observation
that the crystals of racemates of both ethyl ester and
methyl ester of benzene-l,Cdiacrylate are isomorphous
with their respective enantiomers. It has been estab-

Chemical Reviews, 1987, Voi. 87, No. 2 443

Photochemical Reactions of Organic Crystals

osOL-Ba(c103)2

MefoH
no

ti
Me

( 2 R , 3s)

x-x

: Alkene

0 . CO2H

x-x

+
@

: CH3

0 ;

Oh

Figure 10. Hydroxylation of tigilic acid in the solid state.

lished that in the racemate of ethyl ester crystallizing


in a chiral system, the (R)-and (S)-sec-butyl groups are
disordered. The most important conclusion that may
be drawn from these observations is that it is possible
to synthesize optically active polymers and dimers from
achiral monomer if and only if it crystallizes in a chiral
space group and also with favorable intermolecular
arrangements.
Although not a photoinduced reaction, the recent
report of the asymmetric induction achieved by a nonchiral reactant on a nonchiral crystal deserves ment i ~ n . ~ It
~@
was recognized that it is possible, under
certain conditions, to achieve stereoselectively a chiral
product from nonchiral reactants by using one surface
of a single crystal as the chiral template. The conditions
to be satisfied are (a) the projection of the crystal
structure on the reacting surface must belong to one of
the plane groups P1, P2,P3, P4, and P6,and (b) the
molecules must be properly aligned with respect to the
symmetry elements in the crystal. In their experiment,
an aqueous solution of barium chlorate and osmium
tetroxide was allowed to react on the (210) and (210)
planes (plane group P2) of large single crystals of tiglic
acid (space group PI). Enantiomeric diol (Figure 10)
of high optical purity (95%) was isolated.
It may be observed that the creation of asymmetric
molecules from achiral molecules in the crystalline state
using the two entirely independent approaches are
fascinating albeit each imposing certain stringent conditions to be taken into account. However, in light of
the far reaching importance of this subject further intensive study is called for.
6. Competing Dimerization Reactions

A single molecule in the solid state can in principle


undergo several different reactions to give more than
one product. Although each one of these processes is
expected to be governed by topochemical principles, it
is important to establish the factors governing the

competition between these reactions. It is well-known


that in solution the competition between reactions is
controlled by rate processes. Several examples of dimerization occurring in competition with topochemical
polymerization and addition to different chromophores
in the same molecule and addition giving several dimers
have been reported. Although attempts have been
made by Schmidt, the various factors controlling the
competition between reactions in the solid state have
not yet been firmly established. This is one area where
additional input is necessary.
Irradiation of 2-furylacrylic acid in the solid state
gives 4040% dimer and an equal yield of low molecular
weight polymer.32 In the case of 2-thienylacrylic acid
very small amounts of a polymer accompanies dimerization. It has been established through crystallographic
investigations that packing arrangements in these two
olefins control reaction types, namely, dimerization and
polymerization. Schmidt attributes the small yield of
oligomer in the case of 2-thienylacrylic acid, in spite of
a favorable arrangement, to kinetic factors.
Photodimerization of dienes presents once again an
interesting s i t u a t i ~ n trans,trans-Muconic
.~~~~~
acid, its
monomethyl ester, and cis,cis isomer, all of which
crystallize in cells with shortest axes of 4 A, react in the
solid state upon photolysis. A t room temperature all
give vinylsubstituted cyclobutanes of symmetry m.
The all trans acid and its monomethyl ester give, in
addition to dimers, various amounts of oligomers.
Monomethyl ester, interestingly, gives two dimers upon
photolysis in the solid state (Scheme 4). Although the
formation of all the products has been rationalized on
the basis of the packing arrangement, no conclusions
have been made regarding the factors controlling their
competition. It is important to realize that Schmidt has
indeed attempted to seek an answer to this problem
during their studies on butadiene derivatives in the
solid state.34 Solid penta-1,3-diene-l-carboxylic
acid,
penta-1,3-diene-l-carboxamide,
buta-1,3-diene-1,4-dicarbonitrile, styrylacrylic acid, its methyl ester, and
amide all photodimerize on irradiation. The products
are shown in Scheme 4. The cyclobutane derivatives
obtained from these monomers are exactly accounted
for by the topochemical rules (for the details the readers
are referred to the original papers). The example of
trans,trans-mucondinitrileis noteworthy from the topochemical and mechanistic points of view. The monomers are geometrically related in such a way that the
distances between centers of nonequivalent and
equivalent double bonds are 3.64 and 3.96 A respectively. Reaction across the former leads to centrosymmetric products and to mirror symmetric cyclobutanes
across the latter. Whether the predominent dimer always results from the shorter of the two different double
bond contacts could not be confirmed.
Recent findings on etretinate ( 1)35 and 7-chlorocoumarin23are noteworthy in this context (Schemes 5
and 7). Etretinate gives two dimers resulting from
contacts between CI3-Cl4
and Cl<-C12/
and C13-C14 and
Cl0-C9
(-3.8 and -4.4 A, respectively). Interestingly,
the yield of the products correlates with the distance
of double bond separation; shorter distance favors dimerization. An observation on 7-chlorocoumarin suggests that such a correlation may not be valid under all
circumstances. Irradiation of crystalline 7-chlorocoumarin yielded a single dimer (syn head-head). The

444 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

i t

-3

if"

Figure 11. Packing arrangements of 2,5-dimethylbenzoquinone(a) and 2,6-dimethylbenzoquinone(b).

TABLE 4. Center-to-CenterContact Distance (<4.3 8,) of

SCHEME 13

0+
rl

No reaction

0
2L

J3-+
1 I

NO r e a c t i o n

packing arrangement reveals that the two potentially


reactive 7-chlorocoumarin molecules are separated by
4.45 A, this being the repeat along the maxis. Further,
the centrosymmetrically related double bonds are
closer, the center-to-center distance between them being
4.12 A. Translationally related coumarins are expected
to give the syn head-head dimer and centrosymmetrically related coumarins are expected to give the anti
head-tail dimer on UV excitation. The absence of reaction between centrosymmetrically related monomers
in spite of the closer distance (4.12 A) could be attributed to the influence of nearest neighbors, which is
discussed at length in Section VI.
The power of topochemical principles is indeed impressive when they are applied to the solid-state pho-

Double Bonds (C==C,C=O) and Corresponding Dimers


contact
compd
double bonds
distance, 8,
3.49
2,5-dimethyl-1,4- (1-3) &44-5) Alw
benzoquinone"
(1-3) Booo-(4-5) BlW
3.62
(1-3) B,45'-4)
Blm
3.75
(2-3) &.**(2-3) Aim
4.01
(2-3) BW.es(2-3) BlW
4.01
Aim
(2-3) &*..(2'-3')
4.03
2,6-dime~hyl-1,4- (3-5) &41-2)
bl0
3.66
benzoquinonea
(7-6) &-(8-9)
bl0
3.70
3.85
(9-8) A01041-2) A010
3.88
(7-6) &45-3) A010
3.98
(4-8) &49-8) A010
3.98
(3-5) &43-5) A010
4.12
(7-6) &...(9-8)
Bm1
4.12
(3-5) &..*(7-6) A101
4.24
(3-5) &*.*(7-6) A111
aSee Figure 11 for atom numbering

tochemistry of quinones. As illustrated in Scheme 13,


quinones 24 and 25 are photoinert in the solid state
whereas 26,27, and 28 give a large number of products
resulting from dimerization.67@Addition of the C=C
double bond to both C = C and C=O is observed. The
absence of short parallel contacts in 24 and 25 may
account for the inability of these two quinones to yield
dimers. The structures of the dimers (cyclobutanes and
oxetanes) resulting from 26 and 27 can be related to the
packing geometry in the monomer crystals, of parallel
double bonds (C=C, C=O) with center-to-center distances up to 4.3 A. The packing arrangements for
2,5-dimethylbenzoquinone and 2,6-dimethylbenzoquinone and the short contacts between reactive centers
are shown in Figure 11 and Table 4, respectively. Once
again no correlation could be drawn between the center-to-center distance and the yield of various products.
Thus it is clear from the examples presented above
that when there is more than one possible reaction in
the solid state, all can occur if they are topochemically
allowed. However, at this stage, one cannot make an
apriori prediction regarding the extent of feasibility
based on the geometrical factors alone. Obviously the
surroundings should be among the dominant factors in
curtailing the reactivity and controlling competing reactions.

Chemical Reviews, 1987, Vol. 87, No. 2 445

Photochemical Reactions of Organic Crystals

SCHEME 16

SCHEME 14

R E 0COCH-j
h3

/=/R

solid

h3

solid
R

21

Ph Ph

R :CN, COOCH3
SCHEME 15

h3

solid
h3

solid

SCHEME 17
LH3

& h 9

h3

/JPh
solid *
NO2

Ph

No2

N\

NO2

No2

GiiP

%+$$
0

N/

3 : 1

Ph

Ph

solid
h3

Me<

[76,77]
Ph 'N

R = COOMo

Me

Ph

7. Miscellaneous Dimerization Reactions

As mentioned earlier there are a large number of


examples of dimerization of molecules in the solid state.
Most of them do not contain crystallographic details,
often these form part of other investigations. We have
summarized most of them in the farm of schemes in the
hope that some of these systems would attract considerable attention for in-depth studies. Dimerization of
olefins which generally undergo isomerization in solution are summarized in Schemes 14 and 15. Similar
dimerization of nonisomerizable olefins are summarized
in Scheme 16. Several natural products and more
complex olefinic systems are also reported to dimerize
in the solid state (Scheme 17). Intramolecular cycloadditions are known (Scheme 18). Dimerization of
dienes and other cumulenes have also been reported
(Scheme 19). Some of these molecules have attracted
attdntion in connection with photochromism. Known
[4 + 41 dimerization reactions are summarized in

X = H, CI, Br

Q
X

"W
R: COOH

No dimerization

Scheme 20. Dimerizations of imines have also been


reported, and the products are dependent on the substitution (Scheme 21).lo2J03 In addition to these, dimerizations of a few other molecules are also
k~own~104-111

446 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

SCHEME 18

SCHEME 20

h3
solid

[Ref. 91,921

:--$

[Ref. 961

Ph
Ph 'Me

@@
'

Ph

9,

h3
solid

solid

SCHEME 21

'0

'0

SCHEME 19
COOMe

[Ref. 9&]
0

')

(ph
Ph

ph

h3

solid-

PhH:Ph

pv

h3

solid

[95,96]

Ph

Ph

H
dOPh

COPh

Me0

H
H

Nakanishi et al. have reported an unusual phenomenon, namely solvent incorporation during photodimerization of a few cinnamic a ~ i d s . ' ~ ~They
- ' ~ ~found
that a water molecule is incorporated into the dimer
when p-formylcinnamic acid was irradiated as suspensions in water or in a humid atmosphere. Similar incorporation of hydrocarbon solvents into the dimer was
observed when these crystals were irradiated as suspensions in hexane etc. The mechanism of this reaction
is not clear. Although the authors claim that the dimerization does not originate from the dissolved cinnamic acids, in the solvent used for suspension, it needs
further attention.

I I I . Topoehemical Postulate and Unhoiecular


Transformations

The correlation of solid state chemical reactivity with

M e
H
Me0
H
H

H
H
H
M

H
H
e

H
H
But

X-ray crystal structure data has provided a valuable


insight into dimerization reactions. It is now well established that the intermolecular arrangements play an
important role in controlling solid-state bimolecular
reactions. On the other hand, for unimolecular reactions such as intramolecular hydrogen abstraction,
electrocyclization, and fragmentation reactions the intermolecular arrangement is expected to play only a
secondary role. For such reactions the intramolecular
geometrical considerations play a decisive role in controlling the course of the reaction. In other words, the
conformation adopted by the molecule in the solid state
will determine the reactivity pattern. In this section
we highlight the role of packing arrangements in controlling the solid-state behavior of organic molecules
with respect to unimolecular reactions.

Chemical Reviews, 1987, Vol. 87, No. 2 447

Photochemical Reactions of Organic Crystals

SCHEME 23

SCHEME 22
solid and
solution

+pI h m L o

hP h 3

ph@
Ph

solution only

OH

Ph

34 ( R :

35 ( R :

( R : H , R : Me1
( R : R: HI
(R:Me,R:
H)

Rr M e )
R: H I
H,

( R : Me, Rr H )

30 0
-

6H
solution
(Conlormotional
change I
+

H3E

-31

H3i

Solid ond Solution

&+&$
+&
OH

38
-

solution

32

p-H

>

by Oxygen

NO REACTION

33

A. Intramolecular Hydrogen Abstractlon


Reactlons

Scheffer, Trotter, and co-workers have recently elucidated how the ground-state conformations influence
the excited-state behavior of tetrahydronaphthoquinones and their derivatives in the solid state.l15
These studies are concerned with the solid-state
structure-reactivity relationships in a class of organic
reactions not yet well examined, namely, those involving
intramolecular photochemical hydrogen abstractions.
Following these pioneering contributions, work related
to intramolecular hydrogen abstraction in the case of
arylalkyl ketones and aromatic nitro compounds have
appeared in the literature. In general, all these studies
are aimed at answering the following questions about
the factors which influence photochemical intramolecular hydrogen abstraction: (1)Over what distances can
abstraction occur? (2) What is the preferred geometry
for abstraction? (3) Can abstraction be facilitated
relative to the competing processes by freezing a reactant molecule in a particular conformation in the solid
state? (4) Will the products of such reactions in the
solid state differ in type or amount from those obtained
in solution due to crystal packing? A knowledge of such
information is expected to pave the way towards
crystal engineering in unimolecular reactions.
1. Tetrahydronaphthoquinones

Photolysis of cis-4a,5,8,8a-tetrahydro-1,4-naphthoquinone derivatives yields a plethora of products. Im-

SCHEME 24

@
$
a
*
HP

bonding
>6-*

Hp-0

!!

H-0

It
bonding and
heionization*

pressive differences in the nature and yield of the


products were noticed between the solution and solid
state irradiations and these are summarized in Schemes
22 and 23. The mechanism of formation of these
products can be understood on the basis of three primary reactions namely, /3-H abstraction by the carbonyl
oxygen, y-H abstraction by the enone double bond, and
inter- and intramolecular cycloadditions.
For tetrahydronaphthoquinones 29-31 photolysis
generates enone alcohols in the solid state and in solution, and for 29 an additional photoproduct is isolated
from the solution irradiation.ll6Jl7 The formation of
these products can be rationalized on the basis of ,f3hydrogen abstraction by the carbonyl oxygen to form
a biradical (Scheme 24). According to X-ray crystal
structure analyses of 29-31, H,is in a favorable position
to be abstracted by the carbonyl oxygen attached to C1.
Table 5 lists the H, to oxygen distances for these com-

448 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

TABLE 5. Parameters Relevant to p and y-Hydrogen Abstraction by Oxygen in the Tetrahydronaphthoquinones

(a) compdn$c
29
30
31
32
33
ideal values
(b) camp@
32
33
ideal values

Hp.0, A
2.46
2.57
2.58
2.47
2.26
2.12
He-C, A
2.80
2.66
2.90
-____.__.__
Ha-0, A
2.38

(c) compdbtc
34

__

7,

HO-0, A

a See Scheme 22 for structures of compounds.


parameters.

An, derr

C(l)***C(6).
8,

81.3
80.7
83.6
85.1
86.4

3.51
3.46
3.38
3.35
3.33

90

3.40

rC,deg

4, deg

C{3)...C(5), A

52
50

72.6
74.4
90

4, deg

3.17
3.17
3.40
intermolecular
C2-C3 double-bond
separation

101.4
80.8
79.5

4.04
5.21
3.76

45

2.49
2.42

35

dea
3
5
8
0
1
0

_
I

TO,

deg

15
4
3

See Scheme 23 for structures of compounds. See Figure 12 for definition of geometrical

SCHEME 25

bonding

Sum of van der Waals rodii : 0


C

Ideal

To

+H
+C

2.72
= 3.LO A

'0

Ideo1 A 0 = 90'

Figure 12. Definition of angles T~ and during intramolecular


hydrogen abstraction by naphthoquinone.

pounds; they range from 2.26 to 2.58 A. Also included


in Table 5 are the values of 7, and A0.7, is the angle
subtended by the oxygen to H, vector and its projection
on the plane of the C1 carbonyl group. A, is defined as
the angle formed between the carbonyl carbon, the
carbonyl oxygen, and the hydrogen being abstracted
(Figure 12). The ideal geometry for H, abstraction by
the carbonyl oxygen (na*) would have a short H,-O
distance, with the C(8)-H bond in the plane of the
carbonyl group, with C1=O-H = 90". Perusal of Table
5 reveals that the H,-O distances are shorter than the
van der Waals distance of 2.72 A, 7, angles are 3-8" and
A, angles 80.7-83.6', close to the ideal values of Oo and
90, respectively. The biradical formed by the hydrogen
abstraction process is expected to have a conformation
very similar to that of the ground-state precursor, which
is favorable for bonding of C1 and C6 to yield the final
product. Therefore, X-ray structural analyses indicate
that the molecular geometries are favorable for H, abstraction by the oxygen, and subsequent C1-C6 bond
formation to yield the enone alcohol. The formation
of an additional solution product in the case of 29 is
topochemically forbidden in the solid state, since the
conformational changes required for its formation are
prevented by the constraints of the crystal lattice
(Scheme 24).
Naphthoquinones 32 and 33 give similar enone alcohols and in addition cyclobutanone products (Scheme
22). The proposed mechanism for the formation of

Hr

cyclobutanone products involves y-hydrogen abstraction by the enone carbon (Scheme 25) followed by C3-C5
bond formation. The geometrical parameters required
for the mechanistic analyses are given in Table 5. It
is evident that the H,--C2 and C3-C5 distances are less
than the van der Waals distances of 2.90 and 3.40 A and
the 7, and A, are not far from ideal. Thus, the conformations of the molecules 32, and 33 are well suited
for the formation of the cyclobutanone products. The
lack of cyclobutanone formation in the photochemistry
of 29-31 may not be solely due to the unfavorable
geometrical parameters for the hydrogen abstraction by
carbon. The differences in the H-C2 distances are
small (29, 2.97; 30,3.09; 31, 2.86 A; the sum of van der
Waals radii is 2.90 A). The differences in the nature
of the reactive excited state between 29-31 and 32 and
33 may also contribute towards this unusual behavior.
The above studies have thus established that the distances over which abstraction can occur range from
2.26-2.58 A for abstraction by oxygen and 2.66-2.89 A
for abstraction by carbon.
Three of the 1,4-naphthoquinones investigated undergo intermolecular ene-dione double bond [2 + 21
dimerization upon photolysis in the solid state. These
are 34-36 (Scheme 23). In solution, all of these undergo
intramolecular hydrogen abstraction. The question
whether the reactivity difference was due primarily to
a solid-state conformation which is unfavorable for intramolecular hydrogen abstraction or to a particularly
favorable intermolecular crystal packing arrangement
for dimerization has once again been resolved through
X-ray structural studies. Overall, the compounds 34
and 35 possessed the same "twist" conformation common to 29-33. It is obvious from Table 5 that the
geometrical parameters for hydrogen abstraction by
oxygen in the case of 34 and 35 are well within the
guidelines discussed above. However, the major crystallographic difference between the substrates 34 and
35 on one hand and 29-33 on the other was the presence

Chemical Reviews, 1987, Vol. 87, No. 2 449

Photochemical Reactions of Organic Crystals

SCHEME 28

SCHEME 26

Me

*e*-

uii

MoJor solution
photoproduct

&

Solid stoto
conformation

Solid state
photoproduct

"

I
Prolorred when A =CH3
I

c "
i
1 [2 t 23 h,3
cycloaddition

44

"H

:H

8
0

\\

x :OH

Y@

B
H

1 Clorure

No reaction

HO H 0

&

Preferred when

2 Hemiacetel formtion

Closure

&
4E
4F

B u

:&
0

the solid state need further study. For example, while


the naphthaquinones 24-3 1 have molecular structures
favorable for both P-H abstraction by the carbonyl
oxygens and y-H abstraction by the enone double
bonds, only the former occurs. While the naphthoquinones 34 and 35 have arrangements conducive for
both intermolecular dimerization and intramolecular
hydrogen abstraction, only the dimerization takes place.
Furthermore, 39 is inert although intramolecular hydrogen abstraction is expected according to X-ray
structural analysis. Two questions need to be answered
under these circumstances: (a) whether the nature of
the reactive excited state (na* or aa*)is altered by the
substituent and (b) whether the arrangement of molecules immediately surrounding the reactive center play
a significant role in limiting the reaction.

2. TetrahydronaphthoquinolslZ0 -lZ2

of a particularly close and parallel approach of the


C2-C double bond of the neighboring molecules (34,
4.04
35, 3.76 A). Thus, it is clear that the deciding
factor favoring the solid state photodimerization over
the intramolecular hydrogen abstraction is the close and
geometrically favorable approach of the reacting double
bonds.
An example is also available where intramolecular
cycloaddition occurs to yield an oxetane.'17 In this case
(38), the crystal structure neither favors the intramolecular hydrogen abstraction nor the intermolecular
dimerization. Under these conditions the geometrically
feasible but least favored oxetane formation occurs.
Finally, the 2,3,6,7-tetramethyl derivative 39 is unusual
in that no reaction occurs upon irradiation in the solid
state, while in solution the usual photoproducts due to
the P-hydrogen abstraction by oxygen are obtained.
The lack of reactivity in the solid state in spite of the
favorable geometrical parameters is indeed intriguing.
Very recently, Mandelbaum1lBet al. have reported the
photorearrangement of bicyclohept-l-en-l-yl-p-benzoquinone in the solid state. The product of the rearrangement and the proposed mechanism are illustrated
in Scheme 26.
The factors governing the hierarchy of reactions in

1;

Irradiation of naphthoquinols in solution yields


products derived from the intramolecular [2 + 21 cycloaddition (Scheme 27). On the other hand, entirely
different products are obtained upon photolysis in the
solid state. Three different types of photoproducts are
obtained (Scheme 27). These derive from hydrogen
abstractions by the enone carbon and the carbonyl
oxygen similar to the tetrahydroquinones discussed
above. X-ray crystal structure studies have led to an
understanding of the mechanism by which these products are formed and also provide an answer to the absence of formation of intramolecular [2 + 21 adducts
in the solid state.
The tetrahydronaphthoquinols can exist, in two lowenergy conformations as illustrated in Scheme 28.
They can be interconverted by ring flipping involving
a higher energy conformer. The X-ray data of 40-46
revealed that these molecules adopt either one of the
low-energy conformations (A or B, Figure 13) in the
solid state and the preferred conformation is determined by the nature of the substituent at C4 and by the
relative configuration at this center. The preferred
conformations for 40-46 are given in Scheme 17. The
X-ray crystal structure data clearly indicate that [ 2 +
21 photocycloaddition should be topochemically forbidden in the solid state, the double bond separation

Ramanswtky snd Venkatesan

450 Chemical Reviews, 1987, Vol. 87, No. 2

TABLE 6. Geometrical Parameters Relevant to t h e Photochemical Reactons of the Tetrahydro-l-naphthoquinols


compd"
H ( 5 ) * 4 ( 3 )A
,
re, deg
Act deg
c2...c,,A
2.72
2.72
[H(8)*4(3)= 2.921
2.81
2.78
2.78
[ H ( 8 ) * 4 ( 3 )= 2.851
2.84
2.84
2.81
[O(l)*.*Hb= 2.491
2.82
[O(l)***Hb
= 2.491

40
41

42
43
44
45

46
47

48

53.2
53.0
[ r o= 491

50.0
51.9
52.0
[ro = 51)

53.5
54.1
55.7
[ r o = 0.6'1

56.7
= 4.0'1

[io

78.5
78.5
[Ao = 75.21
78.3
17.5
77.5
[A, = 71.61
79.0
79.7

3.30
3.30
[C(2)*4(8)= 3.231
3.35
3.30
3.30
[C(2)-C(8) = 3.171
2.84
3.42
3.39
[C(l)-C(S) = 3.401
3.35
[C(l)***C(S)
= 3.411

aSee Schemes 27 and 29 for structures of compounds. *See Figure 11 for definition of geometrical parameters.

CONFORRRTION

'*R"

Figure 14. Definition of geometrical parameters during hydrogen


abstraction by the olefinic carbon in tetrahydronaphthoquinols.

C O N F O R R R T I O N "8"

Figure 13. Stereodrawing of conformation A (a) and conformation B (b) in the case of tetrahydronaphthoquinols.

being >4.4 A and nonparallel. Thus, it is apparent why


molecules 40-46 fail to undergo intramolecular photocycloaddition in the solid state although this is preferred
in solution. In solution, where the conformational
equilibrium is facile, rapid [2 + 21 photocycloaddition
occurs via the minor higher energy conformers.
The majority of the solid state photoreactions of
tetrahydronaphthoquinols proceed via an initial hydrogen abstraction by the enone carbon C3. The hydrogen that is abstracted is from either C5 or Cs. The
resulting biradicals then collapse by the formation of
either the C2-C5 or the C2-Cs bond (Scheme 28). Interestingly, 4a-01 or anti series (4a-01 anti or trans to
the 4p bridgehead substituents) which crystallize in
conformation A give products derived from the C3-H5
abstraction while 4p-01 or syn series which crystallize
in conformation B yield products via the C3-H8 abstraction (Scheme 27). These are indeed expected on
the basis of the closeness of the reactive centers in these
conformations. Molecules crystallizing in conformation
A are well suited to undergo one more reaction, namely,
H, abstraction by the carbonyl oxygen similar to the

P-H abstraction by the carbonyl in the case of the


naphthoquinones. The geometries for such a process
are favorable for all the members of the anti series.
Table 6 lists the relevant geometrical parameters for
hydrogen abstraction for all the nine compounds. The
important parameters are similar to those discussed for
, and C24!, (Figure
the naphthoquinones: H-C3, T ~4,
14). For the anti series, the H5-C3 distances are well
within the van der Waals separation of 2.90 A and are
in the range 2.72-2.84 A. 7,and 4 are reasonably close
to the ideal values of 90, respectively, 7, N 52-57O and
A, N 77.5-82.2'. Furthermore, the C2*.C5distances are
close to the normal van der Waals separation. Thus,
the formation of products via C3-H5 abstraction is topochemically expected indeed.
A striking aspect of the solid state photobehavior of
the anti series is that only two members undergo carbonyl-hydrogen abstraction, although all would be
expected to do so based on the geometrical factors
(Scheme 29). This clearly illustrates that the examination of the crystal structure parameters alone will not
provide an answer to all the reactivity problems. Based
on the temperature-dependent photochemistry of 47,
the difference in behavior between 47,48, and the other
members of the anti series is attributed to the difference
in nature of the reactive excited state.123 It is suggested
that the carbon-hydrogen abstraction occurs from a
TT*~
state and the oxygewhydrogen abstraction from
a n ~ * The
~ . solid-state results are interpreted as being
due to competing reactions from the
and T T * ~
states, probably existing in equilibrium. Different activation energies for hydrogen abstractions from these
states are suggested as the source of the product ratio
changes with temperature. It is further noted that
methyl substitution on the enone double bond stabilizes
the 7 ~ 7 r *state
~
relative to the n ~ state.
* ~ The question
whether the surroundings influence the product selectivity is yet to be examined.

Chemlcal Reviews, 1987, Vol. 87, No. 2 451

Photochemical Reactions of Organic Crystals

9 &

SCHEME 30-

SCHEME 29
R

+
-

transfer of
Ha to C (3)-

solution
h3

--.%
0

om
Conformation A

47
48

R=H
R : CH3

bonding

h 3 solution

49 A
Two examples of the 4P-01 or syn series crystallizing
in conformation B have been reported.121 Of these one,
namely, 41 does not react in the solid state. It is interestin to note that in this case, the C p H 8 distance
of 2.92 is just outside the limit of the van der Waals
sum (2.90 A). The absence of reaction in 41 is ascribed
to this long hydrogen abstraction distance and the steric
compression (to be discussed in section VI) which would
accompany abstraction. The molecule 44 upon photolysis in the solid state yields a product resulting from
the C3--H8abstraction (Scheme 27). The X-ray structural studies also support this observation. The appropriate geometrical parameters are favorable (2.85 A;
7,50; A, 71.6 A; C3***Cs,
3.17 A).
The studies of the solid-state photobehavior of tetrahydronaphthoquinones and quinols represent a very
fruitful combination of the techniques of X-ray crystallography and organic photochemistry. In addition
to providing information regarding the required geometries for hydrogen abstraction by the carbonyl ( n ~ * ~ )
and olefinic bond of the enone ( T T * ~ )this
,
study has
clearly brought out the utility of controlling the conformation of molecules in achieving selectivity in organic reactions. The concept of crystal engineering
in these systems means the ability to predict the conformation of a prescribed molecule in the solid state.
This has become possible and a few successful examples
are also r e p ~ r t e d . l ~ ~ J ~ ~

3.

p, y -Unsaturated Ketones125

Yet another elegant example of crystal lattice effects


on organic reactions has been reported by Scheffer and
Trotter. As shown in Scheme 30, the &y-unsaturated
ketone 50 yields different products in solution and in
the solid state upon photolysis. The X-ray structural
data suggest that the formation of the solid-state
product can arise either from C,-hydrogen abstraction
by oxygen or through C5-C3bonding. The geometrical
parameters for hydrogen abstraction (oxygen-hydrogen

distance, 2.64 A; T , 28.5; A, 84.7) are very similar to


the values discussed earlier. Furthermore, the feasibility of formation of the biradical via C3-C5 bonding
is indicated by the fact that the p orbitals at C3 and C5
extend toward one another with an internuclear distance of 3.12 A (van der Waals sum: 3.40 A). Between
the two processes, the latter is suggested to be more
probable. The simplest explanation for the differences
in reactivity, originating from a common 7 r ~ excited
* ~
state, between solution and solid state appears to be
that in the solid state, the crystal lattice effects do not
permit the relatively greater atomic and molecular
motion necessary for the formation of the normally
favored solution photoproduct, the result being an alternate unimolecular process involving much less atomic
and molecular movement predominating.
One of the earliest reported examples of rearrangement in the solid state is that due to Matsuura and
co-workers on santonin.lm Unfortunately no structural
details are available to correlate the reactivity in the
solid state. In benzene santonin yields lumisantonin
upon irradiation whereas irradiation of crystals of santonin afforded three dimers, lumisantonin, and a new
photoisomer (Scheme 31). The structure of the dimer
obtained in largest amount has been determined. It has
been speculated that the highly selective transformation
of santonin to 51 in the solid state may be due to the
larger movements associated with the formation of
lumisantonin, the normal solution product.
Recently the photobehavior of two steroidal ketones
of
in the solid state has been r e ~ 0 r t e d . lPhotolysis
~~
&y-unsaturated ketones usually results in a 1,3-acyl
shift to form a new P,y-unsaturated ketone or a 1,Zacyl
shift to give a conjugated cyclopropyl ketone. As illustrated in Scheme 32, irradiation of the 0,y-unsaturated keto steroid 52 in solution as well as in solid state
yields the same products. However, closely analogous

452 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

SCHEME 31

VIEW "2"

1
lumisantonin

-&
1

& @+@

VIEW "2"

SCHEME 32

OH

OH

OH

solution and

solid state

52
-

CH2

Figure 15. Stereoview of the unit cell packing in &y-unsaturated


ketone 52.

VIEW "Y"
53
-

L
NO
solid

reaction

53 reacts only in solution and is inert in the solid state


to UV radiation. Based on the reported structures of
52 and 53 the above variation in the behavior of the two
closely similar molecules can be u n d e r s t o ~ d . ~ ~ ~ J ~ ~
Packing arrangements and molecular conformations of
52 and 53 in the crystal are provided in Figures 15 and
16 respectively.
4. Aryl Ketones 130 - 138

r
VIEW "Y"

The light-induced cleavage and cyclization of organic


carbonyl compounds possessing favorably oriented y hydrogens, termed the Norrish type I1 reaction, is one
of the most well-studied and important photoreactions.
The reaction takes place in solution, the vapor phase,
and in various organized media. In spite of enormous
interest in the mechanistic and synthetic aspects of this
process, only a very few reports have been concerned
with the y-hydrogen abstraction reaction in the solid
state. Reported examples along with solution results
for comparison are summarized in Schemes 33-35. Of
these only N,N-dialkyl oxo amides show remarkable
Figure 16. Stereoview of the unit cell packing in P,y-unsaturated
selectivity in the product distribution in the solid state
ketone 53.
in comparison with that in solution.130 In general, all
other reported ketones show no significant variation in
carbonyl chromophore. The geometrical parameters
their behavior between solution and solid state. Howused in the analyses are the 0.-H distance and 7,which
ever, correlation of structure with reactivity in these
have previously been defined in the case of naphthocases has provided vital information regarding the
quinones and naphthoquinols. The data are summageometrical parameters for hydrogen abstraction reacrized in Table 7. In the case of a-cyclohexylacetophenones, although both p- and y-hydrogens are contions.
veniently situated for hydrogen abstraction, products
Detailed crystallographic data analysis on a-cyclohexylacet~phenones~~~
and o-tert-butylben~ophenone~~~resulting from y-hydrogen alone are obtained. In the
case of o-tert-butylbenzophenone,although in principle
have yielded useful information regarding the geometry
nine hydrogens are available for abstraction, only two
required for y-hydrogen abstraction by the nr* excited

Chemical Reviews, 1987, Vol. 87, No. 2 453

Photochemical Reactions of Organic Crystals

TABLE 7. Geometrical Parametersa for t h e Hydrogen Abstraction i n Aryl Ketones


y-Hydrogen
compd
a-cyclohexyl-p-methylacetophenone
a-cyclohexyl-p-chloroacetophenone
a-cyclohexyl-p-methoxyacetophenone

O.-H.
2.60
2.60
2.61

deg
49.6
42.0
42.5

T,,

0-Ha
3.83
3.83
3.82

&Hydrogen
la,deg

O*..H,

38.4
35.7
37.0

2.57
2.59
2.64

den
6.8
12.7

TR,

11.6

See Figure 12 for definition of geometrical parameters.


SCHEME 33

Ar

:p - t o l y l

55+56/54

55/56

3.5
1.1

1.5
1.7

2.7
1.1

1.3
1.4

2.2

2.4

1.2

2.3

acetonitrite
solid state

Ar = p-chlorophenyl
acetonitrite
solid state

Ar = p-methoxyphenyl
acetonitrite
solid state

m e { . 130

2.6 O%:
U

OH

acetonitrite
benzene
solid s t a t e

i r

57/58
2.0
2.6

COOCH)

2
NHCOPh

0.5

Figure 17. Molecular conformation highlighting the orientation


of hydrogens for abstraction by carbonyl group in diketone 62.

SCHEME 34

SCHEME 35
Ph

Ph
59
-

c : R1:Ph.

61
-

in b r n r m e
in mcthonol

94
17

2c

solid state, OC
i n benzenr
$n mrthonol

83

- - -

22

62
58

29
16

g d

~lidst~u,OC
I benrenr
cn methanol

S8

27

20
34

60

R2:R3:R4:Rg:M~

d : R1:Ph, R ~ : R ~ : H I R J , R ~ : - I C H ~ I Y
a : R1 :R? :R 3 s R4 :R5 : M I

[Ref. 13g

78

are geometrically favorable for abstraction. In this case,


it has not been possible to further narrow down the
hydrogen being abstracted. The common features of
all the above examples are that the 0.-H distances are
well within the van der Waals sum of 2.72 A and the
abstracted hydrogens are -30-50 above the nodal
plane of the carbonyl T system (7). In the diketone 62
reported by M o ~ ~four
, ~hydrogens
~ ~ J are
~ suitably
~
located for abstraction by the two carbonyl chromophores (Figure 17). The product isolated can derive
from any one of these abstractions although it has not
been possible to identify which one is kinetically favored.
Thus, it is clear that a certain amount of deviation
from the coplanar (7 = 0) hydrogen abstraction is tolerable in the solid state. It has been pointed out earlier135that coplanar hydrogen abstraction is not a strict
requirement for the type I1 process and deviation as
high as 60 (7)would lower the rate only by a factor of
four. Especially in the solid state wherein the triplet
state may have a longer lifetime, abstraction reactions
from nonideal geometries (probably with lower rates)
can compete with the decay of the triplet. However,
it is to be noted that thus generated diradicals may have
a geometry different from that obtained in solution and
their behavior will be dominated by the lattice effects.

As seen in Scheme 33, in the case of a-cyclohexylacetophenones, there is a small difference between the
solid state and solution irradiations. The reduced
amount of cyclization in the solid state has been ascribed to the crystal lattice restriction to cyclization.
The motion required for the cyclization sweeps the aryl
and hydroxyl groups through a large volume and would
be expected to be topochemically disfavored in the solid
state relative to the least motion pathway required
for cleavage. Thus the small selectivity observed is
attributable to crystal lattice control of the biradical
behavior. Irradiation of a-adamantyl-p-methoxyacetophenone afforded the less hindered cyclobutanol
(57) and the more hindered cyclobutanol58 in a ratio
of 2.6 in benzene and 2.0 in acetonitrile.136 A dramatic
reversal of stereoselectivity was observed in the solid
state. In contrast to the solution results, laser irradia-

454 Chemical Reviews, 1987, Vol. 87, No. 2

SCHEME 36

Ramamurthy and Venkatesan

SCHEME 37

b N \ R Z

solid
OH

OH

tion of a-adamantyl-p-methoxyacetophenone afforded


the more hindered cyclobutanol isomer 58 as the major
product (Scheme 33). Based on the reactant X-ray
crystal structure, it is suggested that the intermediate
1,kbiradical in the solid state is born in and restricted
to a conformation which is ideal for direct closure to the
more hindered product.
An elegant example of such a phenomenon is provided by Nfl-dialkyloxo amides. N,N-Dialkyloxo amides upon irradiation in solution undergo primary yhydrogen abstraction to give the type I1 diradical
(Scheme 36). This diradical undergoes three types of
reactions, namely, cyclization, elimination, and 1,4-hydrogen migration. The products of the solid state
photolysis of a few oxo amides are significantly different
from those of solution phase photolysis (Scheme 34).
Interestingly, ,&lactones are obtained in larger yields
in the solid state. This has been attributed to the
crystal lattice effect. The formation of oxazolidin-4ones via 1,khydrogen migration involves considerably
more molecular motion of the biradical intermediate
than does formation of @-lactams,via cyclization. The
1,4-hydrogen migration of the biradical64 is restricted
to the planar or nearly planar cisoid transition state
(Scheme 36). This would require rotation of the C(0H)-CO bond of 64 since y-hydrogen abstraction can
occur only from the transoid form of a-oxo amides. It
is suggested that the crystal lattice restricts such a
motion thus favoring the formation of P-lactams.

5. Nitroaromatics 13'

Photochemical hydrogen abstraction reactions of


aromatic nitro compounds have long been known, and
scattered reports on the solid-state photobehavior of
nitro compounds have appeared since the beginning of
this ~entury.l~"'*~
Known examples of this process in
the solid state and a common simple mechanism involving y-hydrogen abstraction are summarized in
Schemes 37 and 38. It is clear that the examples
provided contain a CH group in the position ortho to
the nitro group. Many such molecules have been reported to be light-sensitive in the solid state but no
detailed information on the photoproducts is available.
Unfortunately, X-ray data of those ortho alkyl nitro
compounds, known to be light-sensitive in the solid
state, are lacking. Nitromesitylene, for which X-ray
structural details are available, is reported to be insensitive to light in the solid state.lsO The X-ray
structure of trinitrotoluene is known but not the

aNO2
a:","
CH2R

h3

solid'

Ac

OH

SCHEME 38

A c g NAco

structure of the p h o t o p r o d ~ c t . ' ~Therefore,


~
a more
coordinated effort on this problem is desirable.
The solid state photobehavior of a few crowded aromatic nitro compounds has been investigated by
Dopp.152-156y-Hydrogens are not available in any of
these cases and abstraction from the &carbon occurs.
This primary reaction initiates a sequence of events
leading to 3H-hdole l-oxides as the final products. The
proposed mechanism and the molecules that undergo
this unusual rearrangement in the solid state are summarized in Scheme 39. It is noteworthy that molecules
65-67 do not undergo intramolecular hydrogen abstraction from the benzylic methyl groups, a normal
photoreaction of o-nitrotoluenes, to any measurable
extent. Instead hydrogen abstraction from the unactivated &position (tert-butyl group) is clearly preferred.
In this context, recently reported X-ray structural details on 65-68 are timely and immensely useful to draw
conclusions on the photochemical hydrogen abstraction
reactions of aromatic nitro compound^.^^^^^^^

Chemlcal Reviews, 1987, Vol. 87, No. 2 455

Photochemical Reactions of Organic Crystals

TABLE 8. Intramolecular Geometrical Parameters &levant to Hydrogen Abstraction Involving tert -Butyl and Nitro
Croups- for Compound 65'gb
atom pair
0-H, A
AH, deg
Ao, deg
7 (")
AN, de@
100.4 (7)
-46.1 (7)
112 (112)
2.58 (3)
135.7 (24)
10(13)-*H(83)
129.5 (23)
101.4 (7)
10(14)-*H(82)
2.74 (4)
50.8 (7)
108 (109)
10(13)*-H(93)
2.41 (3)
123.6 (22)
67.5 (8)
-105.3 (9)
56 (48)
87.1 (16)
82.9 (6)
72.0 (6)
88 (82)
20(13)-*H(92)
2.90 (3)
106.5 (9)
55 (47)
2.45 (3)
112.7 (21)
65.9 (8)
20(13)-*H(93)
10(14)*-H(103)
2.42 (3)
130.4 (24)
71.5 (8)
98.1 (8)
61 (55)
20(14)-*H(102)
2.57 (4)
105.9 (21)
83.1 (7)
-74.6 (7)
83 (77)
2.68 (3)
20(14)*-H(103)
100.1 (21)
60.7 (7)
-105.0 (8)
59 (48)
3.640 (5) A
2.773 (6) A

3.550 (5) 8,
2.781 (4) A
'See Figure 19 for atom numbering. bEsd's are given in parentheses, ideal value: AH = MOO; &goo;
outside the Darentheses correspond to N-0-HYP = 90 and those inside to N-0.-HYP = 120.

= Oo; AN = B O 0 . 'The values

SCHEME 39

Sum of van der Walls radii : 0

H = 2.72i

+ N = 3.25i

Ideal 70 = '
0
Ideal A 0 = 90'
Ideal

Xp
5
-

A H = 180'

Figure 18. Defiiition of geometrical parameters during hydrogen


abstraction by the nitro group.

02N*No2
65 X Xi NO2

-66 X
67

Figure 19. ORTEP projection of compound 65 with the atom


numbering scheme and thermal ellipsoids at the 50% probability
level. The hydrogen atoms are represented by spheres of arbitrary
size.

Compounds 65-68 offer a unique opportunity to understand the solid-state reactions through structural
and packing considerations. The X-ray structural data
are able to answer the following questions: (1)Why is
there no hydrogen abstraction from the benzylic posi-

: CH3
: COCH3

tion (Le., from the methyl groups)? (ii) Are there any
hydrogens in the tert-butyl group oriented in the appropriate geometry for abstraction by the nitro group?
It is gratifying to note that in these cases the molecular
geometry and packing considerations allow a unique
identification of the hydrogen being abstracted. Since
the details on all the four compounds are similar,results
on one of them alone are summarized. The ideal geometry for hydrogen abstraction from the n?r* excited
state of the nitro group is expected to be a short 0-.H
distance and angles C-H-.O (AH),N-0-H (Ao), and
T close to 180, 90, and O", respectively. It may be recalled that similar geometrical parameters were used
in the analyses of the photoreactions of enones and
carbonyl-hydrogen abstraction reactions. The definitions of these angles in the case of the nitro group are
illustrated in Figure 18. The general numbering
scheme for the groups involved in the reaction is shown
in Figure 19 and the geometrical parameters used in the
analysis for 65 are summarized in Table 8. Based on
the expectation that short intramolecular O-eH contacts
(less than the van der Waals sum) would favor hydrogen
abstraction, eight possibilities for hydrogen abstraction
by two adjacent nitro groups exist in the case of 65.
However, from a consideration of the values of AH, Ao,

Ramamurthy and Venkatesan

458 Chemical Reviews, 1987, Vol. 87, No. 2

and 7 in Table 8, it seems reasonable to conclude that


two of the eight hydrogens of the tert-butyl group are
better situated for abstraction by the nitro group of 65,
i.e., H(83) by lO(13) or H(82) by lO(14). This reduction
in number based on Structural parameters is remarkable
considering the fact that there are 36 possible modes
of hydrogen abstraction. The geometrical parameters
for hydrogen abstraction from the benzylic methyl
groups are much less favorable than for the two most
probable choices obtained between the tert-butyl group
and nitro chromophore. Thus the geometrical criteria
are helpful in rationalizing the solid state behavior of
65-67. Based on intermolecular interactions a unique
identification of the hydrogen being abstracted was also
possible. In order for the C and N radicals (Scheme 39)
to combine, after the initial hydrogen abstraction, the
tert-butyl group bearing the radical center has to undergo a rotation about the C(l)-C(7)bond either in the
positive or negative direction. Rotation in the positive
direction would bring C(9) in the proximity of N(13)
and rotation in the negative direction would bring C(9)
nearer to N(14). Now the question is whether such a
rotation would be tolerated by the environment in the
crystal lattice and if so in which direction the rotation
about the C(l)-C(7) would be preferred. Based on intermolecular short contacts calculated for the rotation
of the tert-butyl group in both directions up to i3Oo,
rotation in the positive direction is concluded to be less
hindered. Thus abstraction of H(83) by lO(13) followed
by combination of C(9) and N(13) is believed to be the
most favored. The other choice, namely abstraction of
H(82) by lO(14) would not lead to the isolated product
as the coupling of the resulting radicals is prohibited
by the packing. It is obvious that in 65 one cannot
experimentally distinguish between the hydrogens being
abstracted. However, the above approach gains support
in its ability to uniquely identify the hydrogens being
abstracted in the case of 68. It is heartening to note
that conclusions drawn based on the analysis of the
crystal structure of 68 agree with the experimental observation that the nitro group abstracts hydrogen from
the -CH2C1 group of the adjacent chloro-tert-butyl
6. Imines'51 - 173

Photochromism is the salient feature of this class of


molecules. A large number of anils (Schiff bases with
the general formula Ar-CH=N-Ar)
have been reported to be photochromic in the solid
The
most thoroughly studied photochromic solids are the
benzylideneimines, especially the derivatives of salicylaldeh~de.~
Whether
~ ~ ? ~ a~ given
~
ani1 can be photochromic in the crystalline state or not is dependent on
the substituent it carries on the aromatic rings, and in
some cases a variation in the position of a substituent
can prevent photochromism. Even for the same compound, one polymorph may be photochromic while
another is not. Visible light or heat may reverse these
UV-induced color changes.
Early explanations for the mechanism of photochromism in the anils involved aggregation and crystal
lattice interactions. However, investigations of Cohen,
Schmidt, and co-workers showed that isolated molecules
in glassy solutions also exhibit photochromic activity,
thus ruling out such specific solid-state interactions as

SCHEME 40

trans - keto

c i s -keto

enol
enol
enol'+
c&-keto'trans-keto

_h3_

enol*
c& - keto
trans-keto

-bc h -

keto

&

enol

prerequisite to photochromism.166 Although the solidstate studies are important with regard to applications
and understanding the mechanism in the crystals, significant evidence for the intramolecular nature of this
photochromism comes from solution studies. Schmidt
and co-workers first d e m ~ n s t r a t e d l ~
that
J ~ the
~ presence of an o-hydroxyl group is a structural requirement
for the photochromism of anils. In contrast, in glassy
matrix, solutions of anils from other benzaldehydes,
including p-hydroxy and o-methoxy substituted aldehydes are not photochromic. It was postulated that a
six-membered-ring hydrogen-transfer phototautomerism occurs to form a colored quinoid structure. Thus,
in a broad sense this photochromic process involves a
hydrogen-transfer process similar to the ones discussed
earlier in the case of enones, ketones, and nitro compounds.
Based on extensive s t ~ d i e s , ' ~ Schmidt
~ - ' ~ ~ has postulated that the photochromism of anils in the solid
state involves two steps. The colored keto form has
been suggested to be stabilized by a cis-trans isomerization immediately following the hydrogen-transfer
reaction. Formation of the trans-keto form of Nsalicylideneaniline upon excitation in the crystalline
state has recently been confirmed through optical absorption and emission spectro~copy.'~~
Thus the photochromism of anils in the solid state consists of two
reactions, namely, hydrogen migration and geometric
isomerization as illustrated in Scheme 40.
The most extensive crystallographic investigations on
anils have been carried out by Cohen and Schmidt and
more recently by Hadjoudis and C O - W O ~ ~ ~
Their studies indicate that the photochromism is largely
topochemically dominated. They classify two major
crystalline types: the a-type, which is photochromic
and not thermochromic; and the P-type, which is not
photochromic but is thermochromic. Thus in the
crystalline mils of salicylaldehyde,photochromism and
thermochromism are mutually exclusive properties. It
is suggested that the a-type permits the photochemical
formation of the trans-keto structure, whereas the /3type packing prevents this but does permit thermal
formation of the cis-keto structure. In the thermochromic crystal structure molecules are essentially
planar and are packed in stacks of parallel molecules
with an interplanar distance of -3.4 A. In the photochromic crystal, the aniline ring is twisted about the
exocyclic N-C bond by -55O and the molecular packing
is consequently much more open. The packing arrangements of N-(5-chlorosalicylidene)anilineand 2chloro-N-salicylideneaniline,examples of and a type
packing, respectively, are illustrated in Figure 20.
Based on the above packing consideration, it has been
suggested that in the planar molecular structure the
lone pair of the nitrogen does not overlap with the x

~ S . ~ ~

Chemical Reviews, 1987, Vol. 87, No. 2 457

Photochemical Reactions of Organic Crystals

e
a&
CI

ci

CI

CI

CI

CI

B. Fragmentation R e a c t i ~ n s ~ ~ - ~ ~

Figure 20. Stereodrawing of the packing arrangements of N(5-~hlorosalicylidene)aniline


(a) and (a-chloro-N-salicy1idene)aniline (b).
SCHEME 41
0

It

Ph2CH C CH2 Ph

-%
(PhCH2)2 +
-co

Solid s t a t e

P h v R -

Ph2 CH CHZ Ph

Benzene

OR

Ph-C-C-Ph

(Ph2CH)z
1

only

II

OR

Ph-CH-CH-Ph

Ph-CHO

R = Alkyl
Methonol ( R CHJ)
Solid state (degossed)
Solid state (02)

24%

61 %
No r e o c t i o n

PhCOOH (72%)

18 %

PhCOOR (76%)

electrons of the benzene ring whereas in the twisted


structure such overlap is possible. This gives rise to
differences in the basicity of nitrogen, the former, i.e.,
planar form being more basic. The formation of the
cis-keto form can occur readily from the planar as
against the twisted conformation.
Thus, in the solid state an openn crystal structure,
associated with nonplanar molecules and allowing molecular movement, is necessary in order to allow photochromism, whereas in thermochromic crystals the
molecules are planar and closely packed. Such a clear
distinction in the packing arrangements between
thermochromic and photochromic anils allows for
crystal engineering. Two approaches have been suggested in this connection. It has been demonstrated
that the insertion of a nitrogen atom in the 2-position
of the aniline ring of any ani1 which is normally nonplanar (photochromic) yields a planar molecule with
thermochromic ~ r 0 p e r t i e s . lBernstein
~~
and Schmidt
have also demonstrated the use of chloro substitution
in this ~ 0 n n e c t i o n . l ~ ~

A factor of considerable topochemical and practical


importance in unimolecular reactions in crystals is the
fact that the crystal reaction cavity can constrain the
movements of the contents of the cavity and thus give
rise to the specificity of the solid-state reaction. Not
only are fragments constrained to stay close together,
but their rotational motions may also be restricted.
Examples of such an effect are also available in the
literature.
The two examples reported by Quinkert et al. illustrate the restriction brought about by the lattice on
diradicals generated via d e ~ a r b o n y l a t i o n . l ~Photo~J~~
lysis of 1,1,3-triphenylacetonein the crystal gives exclusively a single product, namely, 1,1,2-triphenylethane, while in solution a statistical mixture of products is obtained (Scheme 41). A similar observation
has also been made in the case of 1,3,4-triphenylbutane-2-0ne.l~~Comparative results in the above
systems for solid and solution phases clearly indicate
the translational restriction brought on the radicals by
the crystal structure. The presence of such super cage
effects has also been reported during the photolysis of
benzoin ethers (Scheme 41).177 Benzoin ethers undergo
photochemical a-cleavage to form a benzoyl-benzyl
radical pair which subsequently undergoes free radical
reaction in solution to give benzaldehyde, benzil, and
pinacol ethers. However, when the irradiation of these
crystals was conducted in the absence of oxygen, the
starting material was recovered unchanged even after
prolonged irradiation. Presumably the radical pair
generated in the crystalline phase cannot diffuse apart
and result in recombination. This was indeed demonstrated by photolyzing the above crystals in an oxygen
atmosphere. Under these conditions benzoic acid and
alkyl benzoates, the products resulting from the trapping of the radical pair by oxygen, were obtained in high
yield.
The photochromic behavior reported in the case of
bis(imidazo1es) and P-tetrachloroketodihydronaphthalenes in the crystalline state also result from
cleavage reaction^.^^^'^^ The proposed mechanism is
illustrated in Scheme 42. These are further examples
of the super cagen offered by the crystalline medium.
The radicals generated via cleavage are not free to move
within the crystal structure and strong stereochemical
limitations are imposed on them by the medium.
An example of rotational restriction imposed by the
crystalline medium has also been provided by the investigations of Q ~ i n k e r t . Photolysis
~~
of indanones in
solution results in smooth decarbonylation to give
isomeric benzocyclobutanes. The stereoselectivity observed during the photoelimination of carbon monoxide

458 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

SCHEME 43

VIEW "Y"

MPh
$
Ph

Ph

Ph

ether solution :
cr ys t ol

11 %
9 5 '10

ether solution :
crystol

1 L%

VIEW "Y"

91%
86 %

9 010

SCHEME 44
Ph

Ph

Figure 21. Stereodrawing of the packing arrangement of 2,5dibenzylidenecyclopentanone.

eh

SCHEME 45
H C !

Ph

A
Ph
Ph

Ph

from cis- and trans-l,3-diphenyl-substituted2indanones show a striking increase on going from the
solution phase to the crystal as illustrated in Scheme
43.
One of the most unusual cleavage reactions has been
reported during the photolysis of 2,5-dibenzylidenecyclopentanone.ls4 Photolysis of the above compound
in the crystalline state yielded a spiroheterocyclic compound in addition to the expected dimer (Scheme 44).
The formation of the above product is suggested to
result from the cleavage of the five-memberedring and
subsequent cycloaddition to the C=O bond of a second
molecule. An X-ray structure has recently been reported and the packing arrangement is illustrated in
Figure 21.46 It is noteworthy that the product can, in
principle, be obtained from the starting crystal via the
reaction of nearest neighbors. The distance between
C3-C4 (the bond that is broken) and the carbonyl with
which the diradical reacts is well within the van der
Waals radii. Further, the reacting orbitals are oriented
favorably towards each other.
C. Electrocyclizatlon's5-'s*

Amongst the oldest group of photochromic compounds in the crystalline state is the fulgides which were

60%

LO %

investigated extensively by Stobbe and by Hanel.l*ls7


The crystals of fulgides have initially a yellow to reddish
shade. The effect of light is to deepen the color. All
known photochromic fulgides contain a phenyl group
(Scheme 45). The mechanism of the photoprocess
responsible for this photochromic behavior has been
elucidated by Becker and Santiago1@and established
to be a molecular phenomenon. The primary photochemical step is a photobridging process (electrocyclization) which accounts for the formation of a cyclic
colored structure (Scheme 45). No crystallographic
details are available to further understand this interesting phenomenon.
An example which closely resembles the fulgides is
the cyclization of diphenylmaleonitrile in the crystalline
~ t a t e . ' * ~AJ ~suspension of powdered diphenylmaleonitrile in water, when photolyzed, gave rise to 9,lO-dicyanophenanthrene and 9,10-dihydro-9,10-dicyanophenanthrene (Scheme 45). Diphenylmaleonitrile
shows polymorphism and the photobehavior of the two
polymorphs (prisms and needles) are slightly different.
The effect of polymorphism on the photochemical behavior cannot be fully understood unless the refined
molecular structures of both crystal forms are known.

Chemical Reviews, 1987, Vol. 87, No. 2 459

Photochemical Reactions of Organic Crystals

SCHEME 46
R1

R1

R1

J.

+
t

SCHEME 47

structures of both are available. It is important to note


that only one of these modifications gives rise to a
product upon photolysis while the other is inert to UV
radiation (Scheme 47). The authors have not been able
to provide an explanation as to why the solid form 70
undergoes photolysis but not the form 71. This is partly
due to the fact that it has not been established whether
the rearrangement is an inter or an intramolecular
process. The packing arrangement for the two forms
is shown in Figure 22.

D. Photochemical Oxygen-Transfer
Reactl~ns'~~-~'~

The photoreaction in the solid state occurring from both


the polymorphs, although to different extents, is suggested to be electrocyclization.
A recent study on trans-2,5,2',5'-tetra-tert-butylazoxybenzene by Dopp invokes electrocyclization as a
primary step during its photolysis in the crystalline
state.lgl The products and the proposed mechanism are
shown in Scheme 46. The proposed pathway involves
electrocyclization, fragmentation, and diazonium expulsion. The intermediate 69 has been trapped when
a finely powdered mixture of @-naphthol,sodium bicarbonate, and azoxybenzene was irradiated.
One of the most interesting cyclizations reported is
that of tetraben~oy1ethylene.l~~
Tetrabenzoylethylene
crystallizes in two polymorphic forms, and X-ray

A number of rearrangements are known in which a


transfer of an oxygen atom, originally attached to a
nitrogen, takes place. This group consists of several
potentially useful transformations which have not yet
been fully exploited and investigated.
The transfer of an oxygen atom from the nitro group
to the adjacent unsaturated center has been reported
in several systems both in the solid state and in solut i ~ n . ' ~ Many
* ~ ~ of these, illustrated in Scheme 48,are
complex multistep solid-state reactions. Two mechanisms have been proposed for these reactions and these
can be generally applied to all the transformations
shown in Scheme 48. There is a fundamental difference
in the first step between the two mechanisms (Scheme
49). According to the mechanism proposed by Schmidt
the first step involves the attack of the excited nitro
group on the ortho carbon of the unsaturated substituent whereas de Mayo204postulates a hydrogen abstraction by the photoactivated nitro group with the
formation of a biradical (Scheme 49). The latter process
is thus pictured as a particular case of y-hydrogen abstraction by an excited hetero double bond.
Structure analyses specifically designed to ascertain
the geometry of oxygen transfer have been carried out
on 2-nitrobenzaldehyde and ita halogen derivatives.205i206
The packing arrangement of o-nitrobenzaldehyde is

,
Ramamurthy and Venkatesan

460 Chemical Reviews, 1987, Vol. 87, No. 2

VIEW Y

Figure 23. Stereodrawing of the packing arrangement of o-

VIEW Y

nitrobenzaldehyde.

Figure 22. Stereodrawing of the packing arrangement of the two


(a) and unreactive (b).
forms of tetrabenzoylethylene-reactive
SCHEME 48

azo

solid
h3

ENO
C-OH
II

aNo2
aNo
CH=NR

:lid*

CONHR

R = Phenyl
0

R = N02,

COOH,

COOMe

reaction is still to be fully understood. While the details


of the mechanism of this transformation are obscure,
a conceivable mechanism involving photochemical oxygen transfer from the nitro group to the C(P) in the
ethylenic bridge, a process analogous to the reaction of
o-nitrobenzaldehyde to yield o-nitrosobenzoic acid, is
shown in Scheme 50. The last step involves reaction
of two molecules with water to give indigo and benzoic
acid. It has been shown by Schmidt and his co-workers
that the specific pathway of this complex photochemical
reaction in the solid state is governed by two factors,
namely the molecular conformation and the intermolecular packing.20s
2-Nitrochalcone and those of its derivatives which
are isomorphous with it display the s-trans conformation in the solid state and yield indigos upon irradiation.208This is in keeping with the assumption that this
conformation is necessary for the correct contact geometry between the potentially reactive groups (NO2
and olefinic CB). A better understanding requires full
structural analyses of a large number of 2-nitrochalcone
derivatives.
A formally analogous reaction is the photochemical
rearrangement of azoxybenzenes to azophenones
(Scheme 51).209v210 No structural data are available to
understand the mechanistic details of this solid-state
process. There is abundant literature available on the
solid-state photoreactions of nitro compounds. A
careful scrutiny of them would be valuable.

E. Mlscellaneous Reactions211-221

h3

solid
NO2

shown in Figure 23. On the basis of the reported X-ray


crystal structures it has not been possible to provide
unequivocal support for any one of the mechanisms.
Additional input is required on this important problem.
The solid state photosynthesis of indigo from 2nitrochalcone is one of the most elegant and complicated solid state reactions (Scheme 49). Engler and
Dorant reported207 the transformation of 2-nitrochalcone into indigo by light as early as 1895, but the

The photochemical rearrangement of acetylchloroaminobenzene to p-chloroacetanilide (Scheme 52) has


been the subject of many early investigations.211Pure
dry crystals of acetylchloroaminobenzene,when exposed
to the radiation of a mercury vapor lamp, are converted
rapidly into p-chloroacetanilide. A 90% conversion was
achieved within 8 h. Unfortunately, no structural or
mechanistic details are available for this interesting
rearrangement. Even a speculation of the mechanism
is not possible with the available experimental results.
Interesting transformations have also been reported
to occur during X-ray bombardment of crystals. For
example, crystalline esters of hirsutic acid rearranged
during X-ray data collection (Scheme 52).212 An approximately 1:l mixture of the starting material and the
rearranged product was produced by the irradiation
with X-rays, without disrupting the crystal structure,
and with only minor changes in lattice parameters.

Chemical Reviews, 1987, Vol. 87,

Photochemical Reactions of Organic Crystals

No. 2 481

SCHEME 49

0-

OH

'Schmidt"
(attack on carbon)

0
iI

+
*&Mayo*
(attack on hydrogen)

SCHEME 52

SCHEME 50

@
Ar

h3

solid

c1

.H

. OH
0 %

@q&
0

+q

OH

t
A

II

SCHEME 51

'CHO

CHO

COOH

'COOH

Similarly, when a single crystal of the photooxide of


anthracene was subjected to X-irradiation, it changes
gradually into a mixed single crystal of anthraquinone
and anthrone (Scheme 52).213 This reaction proceeds
via an intermediate stage of disorder, decomposition,
and recrystallization, which can be followed in detail
by means of X-ray diffraction. The "single crystal"
reaction generates - 5 % "free space" in the original
crystal, thus allowing room for the gaseous products of
the reaction to remain within the structure. Caryophyllene nitrosite 72 when irradiated with red light

yields several products.214 The products formed and


the proposed mechanism are given in Scheme 52.
Geometric isomerizations of cis-cinnamic acids and
dibenzoylethylenes have also been investigated.215
Gougoutas has provided a few interesting examples of
solid-state photoreaction^,^^^^^^^ and they are summarized in Scheme 53.

462 Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan

SCHEME 53

SCHEME 55
0

Ar Ar

Ar Ar

@$@@

h3
solid
02

A t Ar

Ar Ar
I

h302

solid

Ph

COOH

Ph CHO

solid

Ph COOH

.OH

X = I,Et, BF4, I 3
0

H0

SCHEME 54

SCHEME 56

PhCOCH-Ph

PhtO tH-Ph

II
--%PhC-0-0-CH-Ph

OR

OR

F. Asymmetrlc Synthesis

Recently Scheffer and his co-workers have reportedn1


the first asymmetric synthesis in the solid state of two
very general classes of unimolecular photorearrangement, namely the di-a-methane reaction and the Norrish Type I1 process. In both instances, very high enantiomeric yields were obtained. Crystals of diester A
(Scheme 54) are dimorphic (space groups Pbca and
P212121). The ketone B (Scheme 54) crystallizes in the
chiral space group P212,21. Irradiations of the Pbca
crystals and solution of A gave optically inactive product C. However, irradiations of P212121 crystals of A
gave C exhibiting average specific rotation of 24.2 f 2.9
(sodium D line). By optically active shift reagent H1
NMR studies, it has been established that the products
are obtained in 100% enantiomeric excess. Similarly
in the case of B, the product cyclobutanol D was obtained in 80% enantiomeric excess while in solution
racemic product resulted. This study clearly illustrates
that the solid state photochemistry holds considerable
interesting results to offer in the coming years.

I V. Structure-Reactlvlty Correlatbns In
Gas-Solld Photoreactions
In a series of papers, Paul and Curtin have provided
a deep insight into the relation between the reactivity
of a molecule in the solid state toward a gas and its
crystal s t r u c t ~ r e However,
. ~ ~ ~ ~ the
~ ~ examples
~
of gas-

PhCOOR

+ Ph COOH

Ph
h9/02
solid
Ph

Ar
Ar)=S

)=o

P
h
?
)7
;(

so2

+s

Ph

Ph

h3/02
solid

R)==O
R

solid reactions reported by them are thermally activated. Gas-olid reactions activated by light have also
been known and all of them involve oxygen as the
gaseous reactant. Most of the reported oxidation reactions in the solid state are summarized in Schemes
55 and 56.
One of the earliest photooxidations to be reported in
the solid state is that of tetramethylrubrene by Hochstrasser (Scheme 55).224 Crystalline tetramethylrubrene undergoes ready oxidation to a colorless
transannular peroxide when illuminated in the presence
of oxygen whereas oxidation of rubrene crystals was
confined to the surface. This difference in reactivity
has been attributed to the possible differences in
permeability of the two crystals to oxygen. Although
this postulate has not yet been verified in rubrene and
tetramethylrubrene through X-ray structural analyses,
the combined photochemical and X-ray studies on a few
other molecules discussed below support the notion that

Chemical Reviews, 1987, Vol. 87, No. 2 463

Photochemical Reactions of Organic Crystals

TABLE 9. Photooxidation of Diary1 Thioketones in the Solid State and Crystal Properties of a Few Selected Thioketones

A : R:R'=H

D
E

117

duration of
irradiation,
days
1
7
15
12
30

F
G
H
I
J
K

201
110
184
258
142
141

30
17
20
30
30
30

crystal mp,
studied "C
A
53
B
75
C
120

CH3

nature of
rxn
in in solid
soln state
crystal data
yes
yes
P2,/n, a = 14.042 A, b = 5.863 A, c = 13.402 A, j3 = 106.4", z = 4
yes
yes
yes
yes
Pbca, a = 7.443 A, b = 32.691 A, c = 11.828 A, z = 8
yes
yes
yes
no
P i , a = 9.810 A, b = 9.635 A, c = 15.015 A, a = 7.11 j3 = 102.30, =
107.76",z = 4
yes
no
RJc,
a = 17.029 A, b = 6.706 A, c = 14.629 A, B = 113.5", z = 4
yes
yes
R12,2,, a = 5.873 A, b = 13.677 A, c = 15.668 A, z = 4
yes
yes
yes
no
yes
no
yes
no

Figure 24. Stereodrawing of the crystal packing highlighting the


presence of a channel in hydrocortisone tert-butylacetate.

permeability of the crystal toward the reacting gas is


essential for efficient gas-solid reaction.
The thermal and/or photochemical oxidation of
crystalline ll-hydroxy steroids to the corresponding
ketones by atmospheric oxygen was reported over 15
years 8 8 0 . ~ ~Most
~ 7 ~recently
~ ~ Byrn et al. have reinvestigated the solid-state light-induced oxidation of the
polymorphs of cortisol tert-butylacetate and their
photooxidation behavior was explained on the basis of
molecular packing in the crystal lattice.227 Cortisol
tert-butylacetate appears to be typical of many steroids
that crystallize in five polymorphs. Two of these forms
are reactive toward oxygen in the presence of light. The
crystal structure of one reactive form has been determined. The crystals belong to the space group B1,
with
the steroid molecules held together by hydrogen

cross section
channel
of the
axis
channel, A*
b
9

3.7

b
a

2.3
8.3

bonding arranged in a helix along the 61-helix axis.


Examination of the molecular packing shows that there
is a channel running through the crystal along the
61-helix axis with a cross sectional area of 3.5 A2 (Figure
24). It has been speculated that the reactivity of this
form towards oxygen is due to the unique packing which
allows penetration of oxygen down the helix axis of the
crystal. Oxygen thus penetrating the crystal oxidizes
the Cll carbon present in this channel.
Photochemical oxidation of 11diary1 thioketones in
the solid state has recently been reported.228 Surprisingly, only six were oxidized to the corresponding carbonyl compound whereas the rest were photostable
(Table 9). However, in solution all were readily oxidized. A comparison of the molecular packing of the
above five thioketones is quite revealing in rationalizing
their photoreactivity in the solid state. For the reactive
thioketones there is a channel along the shortest crystallographic axis with the reactive thiocarbonyl chromophore (C=S) directed towards the channel. In the
case of stable thioketones, the packing arrangement
reveals no such channel in any direction in the unit cell.
The cross sectional areas of the channels are recorded
in Table 9. For the reactive thioketones the channel
cross sectional area is fairly large. The primary step of
oxidation requires an interaction between an excited
thioketone and a ground-state oxygen. For the oxidation to be efficient, oxygen should be able to diffuse into
successive layers of the thioketones. As discussed above
the absence of a channel in 79 and 80 might then be
responsible for their photostability. However, the
presence of the thiocarbonyl chromophore at the crystal
surface is not sufficient for oxidation to occur. It is also

464

Chemical Reviews, 1987, Vol. 87, No. 2

Ramamurthy and Venkatesan


VIEW "Y"

VIEW "2"

Figure 26. Stereodrawing of the packing arrangements of stilbene


(a) and diethylstilbesterol (b).

--.

- C

Figure 25. Packing arrangements in thiobenzophenone (a) and


Michler's thione (b) illustrating the presence and size of channel.
Projection of crystal packing on a plane perpendicular to the
channel axis.

essential that the thiocarbonyl groups be arranged in


a tight fashion so that oxidation of one molecule exposes
another close neighbor to an oxygen molecule. This is
evident from the differences in reactivity among the five
thioketones being considered. Although in all the five
cases the presence of a thiocarbonyl chromophore at the
crystal surface could be identified, the above mentioned
disposition of reactive chromophores is present only in
the reactive thioketones. Typical packing arrangements
for a reactive (thiobenzophenone) and an unreactive
(Michler's thioketone) thione are shown in Figure 25.
The reaction between oxygen and crystals of transstilbene and diethylstilbesterol upon UV irradiation has
been investigated in order to elucidate the role played
by molecular arrangement and defects in the solid
state.229*230
Crystals of trans-stilbene, exposed to UV
radiation in an aerated atmosphere on a microscope
stage, rapidly liquified to give a mixture of benzaldehyde and benzoic acid. The attack of gas is more
rapid on the crystal edges. The attack on diethylstilbesterol crystals is very much slower than on transstilbene. The crystal structure of trans-stilbene is less
tight than that of diethylstilbesterol in which hydrogen
bonds provide strong intermolecular binding. This is

suggested to be responsible for the formation of benzaldehyde faster than p-hydroxypropiophenone. Further
it has been pointed out that these different rates of
oxidation are due to the presence of the ethyl groups
in the a,d positions of the stilbene skeleton; these
groups play the role of a shield for the reaction. Inspection of the molecular packing (Figure 26) of these
two olefins reveals the presence of a channel along the
c-axis with the olefinic bond adjacent to it in the case
of trans-stilbene. Although, diethylstilbesterol has a
similar packing with a channel, the size is smaller. This
may also be expected to contribute toward the difference in reactivity between the two stilbenes. Thus a
similarity in packing and reactivity between steroids,
thiones and stilbenes is evident. The role of structural
defects in the reaction rates seems to be important and
this has been investigated only in the ozonolysis of
stilbenes.
It has been reported recently that crystals of tetrabromofuran undergo photochemical transformation in
the presence of oxygen into the y-lactone 2,3,4,4tetrabromobut-2-en-4olide (Scheme 56).231The initial
step in this transformation is presumed to be the homolysis of an a-C-Brbond.

V. Crysial Englneerlng

Although an understanding of the relationship between reactivity and structure enables one to explain
the product formation and selectivity in many solid
state reactions, these principles are not of much im-

Chemical Reviews, 1987, Voi. 87, No. 2 465

Photochemical Reactions of Organic Crystals

SCHEME 57
h3

solid

%
solid
hS

NO

reoction

NO

reaction

%-

h3

solid

CI

Ph
Ph

Ph
Ph

mediate practical value unless one can engineer a particular polymorphic form possessing the necessary topochemical attributes. There are some obvious difficulties. One of the main problems is the difficulty of
achieving the desired type of crystal structure in any
given case, for the factors that control crystal packing
are not yet well understood. If one had a complete
understanding of the ways in which inter- and intramolecular forces control packing of molecules in crystals
it would become feasible to design template groups of
temporary attachment to functional molecules to guide
photochemically reactive groups into appropriate juxtaposition in crystals. In the absence of such knowledge,
it has been the usual practice to study a series of closely
related compounds, so that a common structural principle can be deduced. The concept of designing molecules so as to guide their choice of crystal structure has
been termed crystal engineering by Schmidt.
The strategy to be adopted may vary from one reaction to the other. Therefore, before embarking on
crystal engineering activity one must have a good
knowledge of the crystal packing, conformation etc.
required for a particular reaction or phenomenon under
consideration. Under the premise that the photodimerization of olefins in the solid state requires an
exact parallel alignment of double bonds the strategies
adopted are outlined in section A-F below.
A. Chloro Substitution

Monochloro substitution and especially dichloro


substitution in aromatic molecules are very effective
devices. Schmidt and his colloborators at the Weizman
Institute have exploited this device to achieve a number
of disparate o b j e ~ t i v e s . ~ ~By
~ - introducing
~~
the dichlorophenyl group into unsaturated systems of the
type Ar-CH=CHX
where X = CHO, COOH,
COOMe, CN, Ph, and Ar = 2,4-, 3,4-, and 2,6-dichlorophenyl, Schmidt et al. succeeded in controlling

the packing geometry with a unit cell having the


shortest axis of about 4 8.
The first indication that dichlorophenyl substitution
would steer the molecule towards ,&packing came
during studies on cinnamic acids.17 While cinnamic
acid crystallizes in both a and p forms, the dichlorocinnamic acids (2,4- and 3,4-, and 2,6-dichloro) crystallize only in the @-form. Since then, a large number
of olefins which do not show any reactivity in the solid
state, have been induced to react through dichloro
substitution. A few examples are provided in Scheme
57. Unfortunately, details of this technique pioneered
and practiced by the group at the Weizmann Institute
are not available. However, following the initial observation by Schmidt, several gr0ups~~1~
have analyzed
the mode of packing of chloro substituted aromatic
compounds. It has become clear through these analyses
that the chlorine atom has a specific role to play in
crystal engineering.
Recent results on the photodimerization of coumarins
in the solid state throw much light on the use of
chlorine as a steering agent for solid-state photodimerizations.22 It is noteworthy that, whereas coumarin itself does not undergo dimerization in the solid
state, all the five chlorocoumarins investigated underwent dimerization in the solid state. Syn head-head
dimers obtained in 6-chloro, 7-chloro, and 4-methyl-7chloro coumarins are the direct consequence of their
@-packingarrangements. Thomas and co-workers have
also employed chloro substitution in the 2-benzyl-5benzylidenecyclopentanone framework.30
Regarding chlorine as a steering group, the generality in the mode of packing in crystal structures containing a chloro group attached to an aromatic ring have
been investigated by us.22 The experimental information for this analysis was taken from the Cambridge
Crystallographic Data Base. Metal complexes and
molecules carrying charges were eliminated from the
analysis. For detailed analysis only structures which
contained Cl-Cl distances <4.2 A were considered. Out
of a total of 132 structures, only 22 did not contain any
Cl-Cl interaction within this limit. This itself speaks
for the success of chloro substitution in crystal engineering. The geometrical parameters used in the
analysis are shown in Figure 27; do is the distance between two chlorines and x is the dihedral angle about
C11..C12 (C1-Cl1-C142). When x = 0 the atoms C1 and
C2are in the cis configuration, whereas x = 180 corresponds to the trans configuration. Figure 27 shows
a plot of the angles O1 vs. O2 (for definition of O see Figure
27). Sixty points lie on the line with O1 + O2 = 180 and
in these cases (x = 0) the arrangement of the molecule
is similar to @-typepacking. Sixty-six points are on the
line with O1 = 02, and this condition simulates the
packing similar to a-type arrangement. In this arrangement C-Cl.-Cl-C are related by a crystallographic
center of inversion. Similarly, in the a-type of packing
the reactive double bonds are related by the above
symmetry. It is noteworthy from Figure 28 which
portrays the plot of N (number of interactions) vs. d
(Cl--Cl distance) that when x = 0 most of the Cl-Cl
distances lie within a narrow range of 3.8-4.0 A whereas
the range is broad (3.5-4.2 A) when x = 180. The
observed smaller width for x = 0 ma:. i e attributed
to the additional interaction between t e close neigh-

466 Chemlcal Reviews, 1987,Vol. 87,No. 2

Ramamurthy and Venkatesan

TABLE 10. Crystallographic Data and Photoreactivities of Members of the Benzylidenecyclopentanone Series
compda
Y
x
a, A
b, A
c, A
space group
dimerization
B-p-C1-BCP (81)
H
B-p-Me-BCP (82)
H
p-Cl-BBCP (83)
p-c1
P-CH,
p-Me-BBCP (84)
p-C1-B-p-Br-BCP (85)
p-c1
P-CH,
p-Me-B-p-Br-BCP (86)
See Scheme 6 for structures of comoounds.

p-c1
P-CH,
H
H
p-Br
p-Br

30.97
31.04
17.18
17.34
17.53

8.50
8.45
10.59
10.68
7.91

18.88

11.21

A
A

601

140120

P2,lC
p2,lC
E I l C

4:

X"0"
j3-Packing

L n +

;2 0 -

VI

;100-

LL

,D 80-

60 -

(1:

k 60-

40

mllc

AROMATIC

40-

no

no
Yes
Yes
no
Yes

HISTOGRAMS CF CI---CI
DISTANCES
VS '10 N U M B E R O F CONTACTS

$-

Pbca
Pbca

180
160-

11.57
11.68
8.80
8.74
11.89
8.29

19

1
,

AROMATIC

- X

-N

180'

OC- Packing
40-

20 -

19

61 ( D E G R E E S

Figure 27. Mode of packing in chlorosubstituted aromatic organic


crystals within C 4.2 A.

bors. One may conclude from the results discussed


above that when there is chlorine substitution, the
chlorine atoms of the neighboring molecules in the
crystal lattice tend to come closer to one another within
a distance of about 4.2 A. Additional interactions arising from the other groups may be expected to steer
the molecule toward either an a or /? type of packing
arrangement. Maybe this is responsible for the dichloro
substitution being more effect,ivethan the monochloro
in driving the molecules towards @-packing.
A convenient method of gauging the magnitudes and
relative importance of weak intermolecular Cl-Cl forces
is to study cases where more than one type of interaction is possible. In this context, a limited survey on the
packing modes of polychloroanilines, phenols, and related compounds has been made.238,239A graphical
analysis has been made of hydrogen bonding patterns
and nonbonded interactions in the crystal structures of
six dichlorophenols. Of the six isomeric phenols, three
crystallize with the @ structure. From their analysis,
the authors have concluded that hydrogen bonding is
the primary factor, but the chlorine atoms are active
in steering the molecule into energetically favorable
packing modes, provided this is possible within the
framework of the energetically stronger hydrogenbonded geometry. In the three @ structures, the chlorine atoms perform the function of bringing stacks
clmer via Cl-Cl interactions, whereas in the three non-@
structures stabilization is achieved by means of strong
Cl-.Cl interactions and minimization of H-H repulsive
forces. The ability for "Cl" to steer the molecules into

CI---CI

DISTANCES

(1)

Figure 28. Histograms of Cl-Cl interactions vs. number of


contacts.

/? structure is felt more in the case of thiophenols and


anilines than in phenols where strong hydrogen bonded
networks are prevalent.
B. Methyl Substttution and the Question of
Isomorphlsm wtth Chioro Derivatives

One of the strategies that has been explored in crystal


engineering operations is the replacement of groups of
equal molecular or atomic volume. In this connection,
the interchangeability of chloro (19.9 A3) and methyl
(23.5 A3) groups has been mentioned in the literature.240
Methyl derivatives frequently crystalize in structures
isomorphous with those of the correspondingly substituted chloro compounds. If this principle holds good,
the promise of methyl substitution as a steering group
is obvious in light of the above discussion on chloro
substitution.
A remarkable observation with respect to chlorinemethyl interchangeability has been made in the 2benzyl-5-benzylidenecyclopentanoneseries.30 In accordance with expectation, compounds 81 and 82 are
isomorphous (Table 10) and both are light stable,
whereas compounds 83 and 84, which are also an isomorphous pair, are photoreactive. The crystallographic
and photochemical properties of compounds 81 to 84
are therefore well behaved so far as the chlorine-methyl
interchangeability is concerned. However, Schmidt has
reported that crystal structures of 0- and p-methyl-

Photochemical Reactions of Organic Crystals


C

Figure 29. Unit cell packing arrangements in 7-acetoxycoumarin


(a) and 6-acetoxycoumarin (b).

cinnamic acids are not related in an obvious manner to


the corresponding chloro acids.17 Further, in the cyclopentanone series itself, compounds 85 and 86 display
a marked difference in reactivity with respect to one
another (Table 10). Further, recent studies on coum a r i n ~demonstrate
~~
clearly that the interchangeability, solely based on the closeness of the size of methyl
and chlorine substituents, cannot be taken to be valid
under all circumstances. The crystallographic and
photochemical properties of the methylcoumarins
turned out to be entirely different from those of the
corresponding chlorocoumarins. While three of the five
chlorocoumarins undergo topochemical dimerization
and pack in a @-typearrangement, none of the corresponding methylcoumarins show topochemical behavior. The validity of the interchangeability or otherwise
of chlorine by methyl may depend upon the importance
of the contribution of the interaction energy of the interchangeable atoms in comparison with the contributions from the remaining atoms.
C. Acetoxy Substitution

Although the significance of the acetoxy group for


bringing the molecules into a suitable geometry in the
solid state for either polymerization or dimerization was
recognized earlier, only recently has a systematic
analysis on the mode of packing in aromatic and aliphatic compounds with acetoxy substitution been carried out.241The results, although encouraging, are not
as clear cut as in the case of chlorine substitution.
Recent studies on coumarins provide some insight
into the role of acetoxy substitution in dimerization
reactions. Of the five acetoxycoumarins investigated,
three underwent dimerization in fairly high yield to the
corresponding syn head-head dimers. The packing
diagram for 7-acetoxy and 6-acetoxycoumarins are
shown in Figure 29. A similar packing arrangement

Chemical Reviews, 1987, Vol. 87, No. 2 487

is expected to be present in photodimerizable crystals


of 4-methyl-7-acetoxycoumarin also. Thus, it is clear
that the acetoxy group plays a strategic role in steering
coumarin rings to pack themselves into a P-type stacked
structure. Packing similar to the above has been reported with 4-(2-carboxyvinyl)-cr-cyanocinnamic
acid
A parallel plane to plane
and its dimethyl ester.242*243
stack is found along the short c-axis (3.956) in which
the molecules overlap completely. But, as to be described below, the analysis based on the Cambridge
Data Base does not show clearly that acetoxy group
substitution ensures P-type packing. From what is
observed in acetoxycoumarins it may be reasonable to
conclude that the nature of the ring system, e.g., coumarin in this case, has also an important role to play
in achieving @-packing.
Interactions involving the overlap of an ester group
of one molecule with the benzene ring of another have
been utilized to steer acrylic acids into packing arrangements suitable for solid-state polymerization.
Similarly, Addadi and Lahav have utilized ester functionality as steering groups during their elegant asymmetric syntheses of chiral dimers and polymers from
benzene-1,4-diacrylate~.~~?~~
In these two cases an attractive interaction between carbonyl and phenyl
groups of adjacent molecules has been proposed to be
responsible for juxtaposing the double bonds at a distance of -4 A.
The above observations suggest that the acetoxy
group is a potential steering group. In order to assess
the generality of this tentative conclusion, a search in
the connectivity file for the acetoxy group has been
performed and some general conclusions have been
arrived at based on 53 cases.241 Out of the 53 cases
examined, 39 had contacts less than 4.2 A between the
acetoxy groups. Of these, 20 structures exhibited dipolar interactions between the carbonyl groups. In six
structures, the arrangement of the acetoxy group is such
that the carbonyl carbon of one molecule is within 4.2
A of the oxygen of the neighboring molecule. In four
cases intermolecular contacts (-3.4 A) are observed
between a methyl group and an oxygen of the carbonyl
group. In only three cases did one find the acetoxy
group oriented in a syn head-head fashion with all
atoms of the group at a distance of -4.0 A (Figure 29).
From the results obtained based on the above analysis,
it appears that the acetoxy group is capable of bringing
potentially reactive centers within 4.2 A, in majority of
the cases through dipolar interactions; however, no
generalizations can be made regarding the preferred
mode of packing.

D. Methyienedioxy Substitution

Very recently, Desiraju et al. have observed that the


presence of the methylenedioxy substituent in a planar
aromatic molecule such as 3,4-(methy1enedioxy)cinnamic acid tends to favor its crystallization in highly
overlapped structures.24 They have noticed that in a
series of (methy1enedioxy)cinnamic acid derivatives,
there is a preference for @-structureover the cx and y.
This has been attributed to the nonbonded interactions
involving oxygen. However, in the case of dimethoxy
derivatives, the @ structure is not expected due to the

488 Chemical Reviews, 1987, Vol. 87, No. 2

SCHEME 58

Ramamurthy and Venkatesan

h3

Ar
A r , Ar = Ph, p - C I P h

R
Ar

Ar

R = CONH2

P-CH3 P h , p - C I Ph

= Ph, p-CH-jPh
C L I

p-tY P O

-1

Ph

Ph

87

es

CN
CN

C N

Ar* Ar

oCOOH ]&A,:
+

COOH

COOH
COOH

A > ~ c o o H
COOH

P-tYPe

repulsive intersheet Ha-H interactions. Such an interaction would become prohibitively large if the sheets
were to be stacked with a 4-A repeat axis. Although the
methylenedioxy group is useful in a few compounds its
generality is yet to be realized. Further, the interactions
responsible for the p stacking are not established.
E. Complexation with Lewis Acids

An additional technique that has been utilized in


crystal engineering is to cocrystallize a,@-unsaturated
ketones with mercuric chloride which itself crystallizes
in a cell of dimensions a = 5.96, b = 12.74, and c = 4.33
A. For example, coumarin which is photostable in the
solid state, forms a 1:l complex with mercuric chloride
which upon photolysis gives the topochemically expected syn head-head dimer. Coumarin in its stable
form crystallizes with a shortest axis of 5.68 A. The
crystals of the complex have a repeat distance of 4.03
A along the c-axis. Similarly, cinnamaldehyde and
benzalacetophenone give crystalline 1:1 complexes with
mercuric chloride having a 4 8, axis.
Recently, Lewis and co-workers reported the photodimerization of Lewis acid complexes of alkyl cin?i, mate
esters with SnC14and BF3.245The exclusive fo
ion
*ysof syn head-tail dimer upon irradiation of r
talline SnC14complex is in accord with the pc
tes
of topochemical control of solid-state dimerizati
he
molecular structure of the 2:l complex (ethylcir
ite:
SnC1,) displays the expected octahedral geomc
rith
Sn at the center of inversion. The distance betvr
the
reactive double bonds in the infinite stacks of e rs is
4.023 A for one symmetry-related pair and 4.12: 4 for
the second symmetry-related pair, well within the range
of values observed for photodimerizable cinnamic acids.
Efficient solid-state photodimerization to yield atruxillate dimers is observed also for the 2:l SnC1,
complexes of the trans isomers of methyl and n-propyl
cinnamates and methyl a-methylcinnamate,
Yet another illustrative example is provided by
(U02)C1, complexes with trans,trans-dibenzylidene-

acetone.246 Praetorius and Kohn reported as early as


1910 that uranyl chloride complexes of dibenzylidene
acetone yields truxillic acid type dimer upon irradiation
as a solid materiaLa7 This is to be contrasted with the
photostability of dibenzylideneacetone in the absence
of urqyl chloride. Light sensitive 2 1 complexes possess
two pairs of ethylenic bonds adjacent to each other
within a distance of 4.1 A. The production of the
truxillic dimer follows directly from the structure of
the complex. The examples thus far reported provide
optimism and certainly encourage further exploration
of this technique.
F. Solid Solution or Mlxed Crystal Formation

A logical extension of the technique of cocrystallization would be to use closely similar molecules to orient
the unreactive olefins in a proper geometry for reaction.
This technique, often termed mixed crystal or solid
solution formation, has been originally investigated by
Schmidt, Cohen, and their co-workersmVm and recently
by others.248,249
Studies by Schmidt on cinnamic acids and cinnamides have shown that one can obtain a solid solution
between either the two reactive or between a reactive
and an unreactive m ~ l e c u l e . In
~ ~all~ of
, ~these
~ ~ cases,
upon irradiation, in addition to obtaining homodimers,
hetero dimers were also obtained. For example, photolysis of solid solutions prepared by melting equimolar
quantities of pairs of P-arylacrylamides (a-forms) produced in every one of the eight examples studied, three
products, two homodimers, identical with the photodimers obtained by solid-state dimerization of the respective solid monomer amides and a hetero dimer,
identified in each case as the pseudocentrosymmetric
dimer (Scheme 58). Very recently mixed-crystal formation between 6-chloro-3,4-(methylenedioxy)cinnamic
acid and 2,4-dichlorocinnamicacid has been reported.247
Both individually crystallize in the @-structureand yield
topochemically expected P-truxinic acids. Mixed
crystals with the ratio 2:l (87 and 88) are obtained from

Chemical Reviews, 1987, Vol. 87, No. 2 469

Photochemical Reactions of Organic Crystals

SCHEME 59

R'

Ar'.,

+ '&%

QR
'

Ar
A r = p-CI-C6H4

Ar'
CH,-C6H4

Ar'

R =COOH

SCHEME 60
A

R + A r b R+
R

Ar'

(01

Figure 30. Molecular conformations of 89 and 90 as found in


their crystals. Neuman projections (top) and computer drawings
(bottom).

SCHEME 61

+
c'=F@
NO

reaction

'r

90
-

'WBr
Dimer

89

CwBr&

"WBr
+

90

N 3 : l

p-chlorocinnamic acid which crystallizes in a @type


structure, and p-methoxycinnamic acid which gives a
light stable y-type, gave two different kinds of mixed
crystals affording, on irradiation, different product
distributions (Scheme 60). It therefore appears that
small quantities of guest molecules can stabilize polymorphic forms which are normally metastable or unavailable.
The above type of study has recently been extended
to crystals composed of solid solutions of various benzylidenecyclopentanones which differ only by a chloro
and methyl inte :hange.248p249The most noteworthy
result was obtai-led in the cocrystals of 89 and 90.
Ketone 89 crystallizes in a photoactive modification
while compound 90 crystallizes in a photostable form
(Scheme 61). The main difference between the forms
lies in the molecular conformation (Figure 30). Yet,
these two compounds form mixed crystals, isostructural
to ketone 89 with the reactive centers in the mixed
structure separated by 3.866 A. Photolysis of these
mixed crystals yielded an unsymmetrical cyclobutane,
the structure of which is yet to be established.
It appears from these results that the strategy of
mixed crystal formation may be successfully used to
incorporate molecules into new and photoactive modifications. Further, it is clear that two compounds
having completely different crystal structures form
mixed crystals primarily as a result of forced changes
imposed upon the molecular packing of one component
by the second component.
Very recently Thomas et al.250have proposed a general method of predicting the packing of the molecules
in crystals. Based on extensive analysis of the crystal
structures of BBCP derivatives, the influence of substituents on solid-state reactivities and modes of molecular packing have been rationalized. They claim that
this approach could lead to engineering of solids to
desired reactivity.

09

Mixed
dimer

ethanol, and photolysis of this solid solution leads to


three mirror symmetric or pseudomirror symmetric
(homo and hetero) dimers. It is of particular interest
to note the behavior of solid solutions from monomers
which by themselves crystallize in a and p types. Two
such pairs were
Cinnamic (a) and p chlorq r.innamic (p) acids; and p-methyl (a)and p chlor(Annamic (p) acids. Each pair produf-edsix dimers, the a and p homodimers of each mo: >mer and
the two (a and 0) heterodimers (Scheme 59). The
results reported above may be most simply interpreted
by assuming a degree of mutual miscibility of each of
the two constituents in the phase of its partner, without
any drastic change in each others crystal structure.
The most useful and interesting situation would be
to induce reactivity of the photostable form through the
formation of mixed crystals of a photostable compound
with a reactive one. This has indeed been achieved. For
example, an equimolar solution of an a-type monomer,
p-methylcinnamic acid, and a y-monomer, p-methoxyG. Unlmolecular Reactions
cinnamic acid, yields three a-dimers, i.e., dimers from
both p-methyl and p-methoxycinnamic a ~ i d s . ~ ~ ~ In
g ~principle,
~~
the reactivity predictions should be
Further, an equimolar solid solution prepared from
easier to make for unimolecular processes since con-

470 Chemical Reviews, 1987, Vol. 87, No. 2

SCHEME 62

&R OH

91
-

&

Ramamurthy and Venkatesan

cH

TRANSIT ION
STATE

PRODUCTS

R
0

92

Proforod when R = C H 3

REACTANTS

93
Proforod w h e n R

:H

0
I

96b
-

H#?

94

H@OH
95
-

formation rather than intermolecular orientation is


often the controlling factor. As opposed to dimerization
reactions, systematic studies on unimolecular transformations in the solid state have been limited. This
is also reflected in crystal engineering investigations.
Such an attempt has been made only in one unimolecular reaction.
It has been established that altering the configuration
at C(4) in analogues of 91 changes the solid-state photoproduct (Schemes 28 and 62).ln Based on systematic
studies it became clear that the bulkier of the two
groups at C(4) adopts the pseudoequatorial position
which in turn determines the preferred conformation
of the molecule in the solid state and hence the photochemistry. This logic led to correct predictions of
products in 92. The solid-state conformation of 92 is
predicted to be A. Product 93 is expected and indeed
obtained in >75% yield in solid state.122
V I . Subtier ASp8CfS of Phoforeactivity in Solids

A. Role of Neighbors in Solid-state Reactions:


Concept of Reaction Cavity

The topochemical postulate states that reactions in


the solid state occur with minimum atomic or molecular
movement. This implies that the reactive centers must
initially be properly juxtaposed in the crystal for the
reaction to occur. Further, it is believed that the photochemical processes in molecular crystals are constrained by lattice forces which are assumed to maintain
electronically excited molecules in structures close to
the positions and orientations they had in their ground
states. This concept of preformation is obviously deficient in at least two respects. Firstly, it does not take
into consideration the importance of nearest neighbors.
The role of the nearest neighbors is reduced to that of
passive spectators during the progress of the reaction.
Secondly, it does not give due consideration to the
changes that may be caused by molecular excitation.
In essence the topochemical principle is an expression

Figure 31. Pictorial representation of reaction cavity concept.

of the mechanical stiffness of the lattice surrounding


the reaction site, allowing only small displacement as
the excited reactants progress towards the products. In
this context, Cohen has advanced the concept of the
reaction cavity which gives due importance to the
presence of nearest neighbor^.^^^^^^ More recently Craig
and his co-workers have emphasized the possible effect
of electronic excitation in a molecular ~ r y s t a 1 . ~ ~ ~ - ~ ~
They have introduced a new concept termed dynamic
preformation. These concepts which are refinements
of original topochemical principles enable one to understand many photoreactions which are apparent violations of original topochemical rules.
Cohen introduced the concept of the reaction cavity
in the crystal as an aid to interpret the course of a
variety of solid-state reactions. This concept has been
extensively utilized by him to understand the geometries of the excimers of polyaromatics in the crysta1.257-263
The molecules taking part in a reaction occupy a certain size and shape in the crystal. This space
occupied by the molecules is the reaction cavity and is
surrounded by other molecules. The atomic movements
constituting the reaction would cause pressure on the
cavity wall which may tend to become distorted. This
distortion, then, involves a large decrease in the number
of attractive forces and/or a large increase in the number of repulsive forces. However, any such distortion
in shape would be restricted by the close-packed environment; as a result only those reactions which involve
minimal change in the external contacts of the reacting
molecules would be energetically feasible. Therefore,
the topochemical postulate can be redefined as
reactions proceeding under lattice control do so with
minimal change or distortion of the surface of the reaction cavity. This concept is valuable in predicting
the course of a reaction when more than one reaction
is topochemically permitted. If more than one product
is topochemically permitted, the preferred one will be
that for which the transition state involves the least
change in shape of the reaction cavity. Consider two
hypothetical reactions giving products B and C from
A (Figure 31). The change in shape of the reaction
cavity, and therefore the change in energy of interaction
with the surroundings, will be different for the two
reactions. While for reaction I the shape of the cavity
remains practically unchanged during reaction, for I1
the transition state is energetically very unfavorable
because of the bulges and voids formed. Thus the
nearest neighbors constrain the system to act by path
I (to give B) rather than by path I1 (Figure 31).

Chemical Reviews, 1987, Vol. 87, No. 2 471

Photochemical Reactions of Organic Crystals

SCHEME 63

Rr. R3

R,1

,R1

Me

98
-

K3

Two elegant examples have become available recently


which illustrate the use of the reaction cavity concept
in predicting the course of solid-state reactions. Enones
of general structure 97, when irradiated in the solid
state undergo one of the two possible photorearrangements (Scheme 63) as discussed in detail in section 111.
Scheffer and his co-workers have reported recently that
in one instance, however, that of enone 98, irradiation
in the solid state lead exclusively to the product resulting from initial hydrogen atom abstraction by the
a-carbon atom, of the enone chromophore (path C,
Scheme 63).391264Hydrogen abstraction by both C, and
C, carbons are favorable in the case of 98 (H-4,: 2.74
and H.-C,: 2.70 A). Thus the preference for C, abstraction is not evident from the geometrical criteria.
This suggests that nearest neighbors control the hydrogen being abstracted. A change in hybridization at
C, and C, from sp2to sp3 is expected to accompany the
hydrogen transfer process and this would force the
methyl groups at these centers into close contact with
certain hydrogen atoms on neighboring molecules and
thus sterically impede the reaction. Computer simulation of the pyramidalization process at C, and C, in
98 (with stationary lattice neighbors) showed no short
contacts (termed steric compression) during pyramidalization at C, and significant steric compression
during the pyramidalization at Cp Figure 32 shows
stereodiagrams of enone 98 before and after pyramidalization at C,. The steric compression accompanying
full 55 pyramidalization is indicated by the dotted lines
and consists of short hydrogen-hydrogen contacts. An
estimate of the steric compression energies accompanying pyramidalization was made using semiempirical
calculations (Allingers MM2 force field program and
Lennard-Jones 6-12 potential function). The steric
compression energy for fully pyramidalized 98 at C,
corresponds to -12 kcal/mol. This, when compared
with zero steric compression during pyramidalization
at C,, accounts for the preference in hydrogen abstraction by C,. Thus, it is clear that consideration of
the unfavorable interaction between the reacting molecules and their stationary neighbors could provide an
insight into the mechanism of solid-state reaction.
A similar observation was made in the case of 7-

Figure 32. Stereodrawings of enone 98 before (a) and after

pyramidalisation at the p-carbon. The steric compression contacts


are shown by dotted lines.
7-CHLORO C O U N A R I N - T R R N S L A T E D - I N I T I A L

ORIENTATION

Figure 33. Stereodrawing of the packing arrangement of 7chlorocoumarin.

chloro~oumarin.~~
Irradiation of crystalline 7-chlorocoumarin yielded a single dimer (70% yield) assigned
the syn head-head configuration. The packing arrangement shown in Figure 33 reveals that there are two
potentially reactive pairs of 7-ChlOrWOuin. One pair
being translationally related has a center-to-center
distance of 4.54 A. Further, the centrosymmetrically
related double bonds are closer, the center-to-center
distance between them being 4.12 A. Translationally
related coumarins are expected to yield syn head-head
dimer whereas the centrosymmetrically related coumarins would give anti head-tail dimer upon excitation.
However, the only dimer obtained on excitation corresponds to the syn head-head dimer. The reason for the
absence of reaction between centrosymmetrically related monomers in spite of a closer distance (4.12 A) is
not immediately obvious. We have estimated265the
energies involved in bringing these two pairs of molecules to a geometry ideally suited for dimerization in
the crystal lattice. It is believed that the best ?r overlap
between the reacting double bonds can be achieved
when 01,02, 03, and O4 (Figure 5 for definition) are 0,90,
90, and O, respectively. One of the assumptions in this

Ramamurthy and Venkatesan

472 Chemical Reviews, 1987, Vol. 87, No. 2

TABLE 11. Results of Lattice Energy Calculations for Reactive and Nonreactive
initial arrangement
initial energy, final energy, rise in energy,
d , A O,, deg 02, deg Os, deg 04, deg
kcal mol-'
kcal mol-'
kcal mol-'
compd

methyl m-brr 2ocinnamate


3.93
4,4,8a-trimetl jl-8afl-carbomethoxy-4a/3,5,8,8a3.79
tetrahydro-1(4H)-naphthalen-l-one
3.83
7-methoxycoumarin
7-chlorocoumarin
pair I (translational)
4.454
4.12
pair I1 (centrosymmetric)
8-methoxycoumarin
4.07
pair I
pair I1
3.87
3.86
methyl 6-isobutyl-2-methyl-4oxocyclohex-2-enecarboxylate
3.83
7-acetoxycoumarin
3.90
6-acetoxycoumarin
3.66
benzylidenebutyrolactone

38.2
0.0

65.2
81.65

104.8
63.5

31.5
0.0

-20.4
-21.5

6705.6
1482.9

6726.0
1514.4

67.5

109.9

66.5

0.5

-37.3

162.8

200.1

0.0
0.0

131.8
127.9

85.3
73.0

0.0
0.0

-17.9
-17.9

159.1
18064.9

177.0
18082.8

0.0
0.0
0.0

122.4
117.4
102.2

63.8
67.7
108.1

0.0
0.0
0.0

-37.2
-37.2
-19.7

7.8
-9.1
10.0

45.0
28.1
29.7

0.0
0.0
0.0

106.4
94.1
109.3

110.3
118.3
77.0

0
0
0

-20.1
-21.2
-18.5

147.6
-12.9
-13.6

167.6
8.3
4.9

For a definition of geometrical Darameters see Figure 5.

For structures of compounds see Scheme 8.

approach is that the first step in dimerization involves


an attempt by both the molecules to reach a full T
orbital overlap. But in these two pairs O2 and O3 deviate
significantly from the ideal value of 90' (translated pair,
82 = 131.8', O3 = 85.3', d = 4.454 A; centrosymmetric
pair, B2 = 127.9'9, O3 = 73', d = 4.12 A). Lattice energy
calculations were carried out using a computer program
(WMIN) developed by Busing.z66 Molecules were
treated as rigid bodies, rotated and/or translated by
choosing orthogonal coordinate system appropriately.
All the symmetry related molecules were subjected to
the motional operation simultaneously. The program
does not allow for rotation of an individual molecule in
the crystal lattice. It is not possible in the program due
to Busing266to keep the molecules surrounding the reacting partners stationary while giving necessary rotations and translations to the reactants. The rise in
energy to achieve the ideal geometry in the crystal
lattice for the translated pair was 177 kcal/mol whereas
for the centrosymmetric pair the energy increase was
much larger (18082.8 kcal/mol). Although the absolute
values based on the above calculations are not meaningful, the relative values are generally useful to explain
the excited-state reactions. It is evident then why the
dimerization is preferred between the translated pair
although the distance criteria alone would lead us to
predict the reaction between the centrosymmetric pair.
The two examples provided, one each from unimolecular and bimolecular photoreactions, illustrate the importance of considering the role of nearest neighbors
in controlling the course of solid-state reactions.
The reaction cavity approach is also helpful in understanding the nonreadivity of a few molecules in spite
of a favorable topochemical arrangement in the solid
state. Compound 2 (Scheme 8), although crystallizing
in a lattice arrangement which is ideal for intermolecular [2 + 21 photodimerization, surprisingly exhibits
complete lack of photoreactivity when irradiated in the
solid state. The potentially reactive double bonds are
oriented in a head-tail fashion and are parallel, directly
one above the other and only slightly offset along the
double bond axis. The geometrical parameters for 2,
useful to identify the orientation of the double bonds,
are = ,'O O2 = 81.65', O3 = 63.5', O4 = ,'O and center-center distance 3.79 A. Scheffer and his co-workers
have a t t r i b ~ t e d ,the
~ ~lack
, ~ ~of~reactivity to the steric
compression (short contacts) developed between the

Figure 34. Stereopacking diagram of enone 2 (see Scheme 8).

reactive molecules and the nearest neighbors when reactive molecules X and X start to move towards one
another. The packing diagram in Figure 34 reveals the
situation. The molecules Y and P act as stationary
impediments to photodimerization. Using semiempirical calculations they have shown that both the
single-motion (one molecule moves and the other is
stationary) and the dual-motion pathways (both the
reactive molecules move toward each other) involve
significant activation energy. We have come to the
same conclusion using lattice energy calculations.265As
elaborated above we have estimated the lattice energy
using the WMIN program for various geometries of the
potentially reactive pairs of molecules in the crystal.
Table 11summarizes the results for a few reactive and
nonreactive olefins. It is clear that for the reactive
olefins the rise in lattice energy to bring the two molecules to the total a orbital overlap is small whereas for
the nonreactive olefin 2 the rise in energy is enormous
(1514 kcal/mol-'). As mentioned earlier the absolute
values are not meaningful although the relative values
could be of some help. Therefore, it is clear that the
nearest neighbors or the surroundings could inhibit the
progress of the reaction even though the reactive
molecules could be suitably geometrically arranged in
the crystal.
Recently, Gavezzotti has explored the role of molecular environment in the crystal towards reactivity and
has generalized that a prerequisite for crystal reactivity
is the availability of free space around. the reaction
site.267 He has developed a computer program which
can be used to perform volume analysis of various kinds
of molecular systems and structured media.%'-= Using
this he has successfully identified the presence of a void
in the crystal structure of azobis(isobutyronitri1e)

Chemical Reviews, 1987, Vol. 87, No. 2 473

Photochemical Reactions of Organic Crystals

SCHEME 64

Ph

Ph
Et-t-N=N+Et
Et

h3

Et

Ph

HC-Et
1

Et

Ph

CH3

Et

>=(

Benzene
Crystal

Ph

>=(

Et

rv

CN

&

>- +

CN

Benzene
Crystal

+Ph#Ph

,u

30 10

<5

O/O

+- +

Et

Et

70 /o

<5

Et

P
h
m

Et Et

CH3

(CH3)ZCHCN

Et

Et Et

~ 7 5 %

25%

CN
t N = N t

O/O

)=C=N+CN

CN CN

4 010
7 a v0

Figure 35. Di (the fraction of the occupied space) map of azobis(isobutyronitri1e) before (a) and after (b) elimination of nitrogen.

(AIBN). The solid state photoproducts of AIBN are


mainly isobutyronitrile and methacrylonitrile (Scheme
64). The Di(the fraction of occupied space in the i-th
elementary volume) map for AIBN shown in Figure 35
is indicative of the packing mode of this compound,
which consists basically of a parallel arrangement of
molecules with large pockets of intermolecular free
space. Gavezzotti proposes that the reactivity of AIBN
crystals is greatly favored by the voids in proximity with
the azo chromophore. These voids provide space for
the incipient N2 molecule and help relax the otherwise
prohibitive internal stress that results from its formation.

B. Photoinduced Lattice Instability: Concept of


Dynamicai Preformatlon
The static concept of preformation or a topochemical
postulate based on ground-state lattice equilibria is
deficient in not taking into account changes caused by
molecular excitation. Craig and his co-workers, based
on extensive theoretical calculations, have drawn attention to another type of preformation, dynamic
preformation which may influence product formation.253-256According to them dynamic topochemical
rules may control in many cases the product formed
and may even determine whether a reaction proceeds
at all. In such a model there is transient preformation
of the reactant pair not evident in the equilibrium local
structure. Thus,Craig suggests that a reaction that may
not be expected based on the ground-state structure
may indeed occur if the dynamic preformation can be
tolerated in the crystal lattice by the nearest neighbors.

L 0Olo
10

5 810
12

We present below a qualitative picture of the model due


to Craig which can be of considerable use in understanding solid-state reactions. This model is expected
to play a significant role in coming years and attract
considerable attention of experimentalists.
Localized electronic excitation of a molecular crystal
is expected to produce a particular type of instability
of the lattice configuration leading to large molecular
displacements. Localized excitation means the existence of an excited molecule which on account of its
altered properties is seen by its neighbors as an impurity. The creation of this impurity molecule introduces a local instability in the lattice configuration
and leads to relaxation. This relaxation process could
involve large displacements from the original lattice
structure and in that sense far from the equilibrium
configuration of the unexcited crystal. This logic has
indeed been found to be valid by Craig and Collins253
for a simple one-dimensional lattice. Based on a simple
monoatomic lattice model they conclude that photoexcitation does produce a local instability in the
lattice structure which is relieved by large amplitude
displacements in more than one possible way.
In three recent papers, Craig and his co-workersw-256
have cited examples where excimers or exciplexes form
on excitation in the solid state but are not preformed
in the ground state. They have shown through extensive theoretical calculations that in these polyaromatics,
a short term lattice instability created via excitation has
the effect of driving one molecule close to a neighbor
thus promoting excimer or exciplex formation. Lattice
energy calculations on excited molecules in crystals of
anthracene, 9-cyanoanthracene, and 9-methylanthracene were carried out by Craig and Mallet254
using a simple center-of-mass attractive excitation potential. Interestingly, excitation of the crystal of 9cyanoanthracene led to the result that for a short period
after excitation, an excited molecule can be displaced
away from its equilibrium position into a asymmetrical
local structure, with the excited molecule closer to one
neighbor in the stack of molecules than to the other.
In such a model there is a transient preformation of an
excimer not evident in the equilibrium local structure.
This is illustrated in Figure 36. It is clear that the
excited molecule moves increasingly off-center towards
its nearest neighbors in the stack with increasing excitation potential parameter. For 9-methylanthracene
there is also a significant movement on excitation

Ramamurthy and Venkatesan

474 Chemical Reviews, 1987, Vol. 87, No. 2


7-flETHOXY

COUflARIN - - I N I T I A L

ORIENTATION

Figure 38. Stereodrawing of the packing arrangement of 7methoxycoumarin.

Figure 36. The structure of 9-cyanoanthracene containing an


excited molecule and allowing the near neighbors to relax their
positions and orientation.

Figure 37. An excited molecule of 9-methylanthracene in a


relaxed environment of nearest neighbors.

(Figure 37). There are similar indications for the mixed


crystal 9-methoxyanthracene in 9-cyanoanthracene as
well as 9-cyano- and 9,lO-dimethylanthracene mixed
crystals. Such findings draw attention to the need in
experiments to look for evidence of dynamical factors
in solid-state reactivity.
On the basis of the above presentation, dynamical
preformation can be summarized as follows: short-term
lattice instability caused by photoexcitation can have
the effect of driving one molecule close to a neighbor

so as to cause a photochemical reaction. If a reaction


proceeds in the short time available before general
lattice relaxation to the equilibrium ground-state
structure it can be said to be dynamically promoted or
preformed. We discuss two cases from the literature
which may indeed be ideal examples of the dynamic
topochemical postulate.
X-ray crystal structure analysis of 7-methoxycoumarin reveals that the reactive double bonds are
rotated by 6 5 O with respect to each other, the centerto-center distance between the double bonds being 3.83
A,24 In spite of this unfavorable arrangement, photodimerization occurs in the crystalline state yielding syn
head-tail dimer as the only product. As seen in Figure
38, the two double bonds, although within the reactive
distance, are not suitably juxtaposed for dimerization.
Having ruled out defects as the possible loci for the
reaction, we have investigated the possible presence of
a certain degree of inherent orientational flexibility of
the molecules in the crystal lattice.21*285
We believe that
radiation energy absorbed by the reacting molecules is
sufficient to allow the molecules to undergo the required
rotation, provided the motion is cooperative and extends through the crystal. In order to investigate the
inherent flexibility that may be available for such a
cooperative motion, lattice energy calculations for the
monomer crystal were carried out. This approach, although different from that of Craig's and lacking its
rigor, yields results that are revealing. Lattice energy
calculations performed using the WMIN program show
that the energy increase to bring the two reactant
molecules to proper orientation is only 200 kcal/mol,
roughly similar to the value obtained in many ideally
oriented pairs (Table l l ) , discussed earlier. In most
crystals, electronic excitation increases the attractive
forces, so that the excited molecule is bound more
tightly to its nearest neighbor. With an increase in
attractive forces between the reactive molecules upon
excitation, one may expect that the motion of the
molecules towards the maximum overlap geometry
would involve less energy than that for the ground-state
molecules. We believe that this is indeed the case with
7-methoxycoumarin. This example when compared
with the photoinertness of methyl m-bromocinnamate
is definitely interesting. In this case, similar to 7methoxycoumarin, the double bonds are not topochemically oriented for dimerization (Figure 4). The
distance between the centers of adjacent double bonds
is 3.93 A, but the double bonds are not parallel. They
make an angle of 2 8 O when projected down the line
joining the centers of the bonds. Based on the behavior

Chemical Reviews, 1987, Vol. 87, No. 2 475

Photochemical Reactions of Organic Crystals

of 7-methoxycoumarin one would expect this molecule


to undergo dimerization in the solid state. But it is
photostable. The reason becomes obvious when one
carries out lattice energy calculations. The energy increase to align the molecules parallel to each other in
a geometry suitable for dimerization is enormous (
6726 kcal/mol) when compared to 7-methoxycoumarin.
Such an increase will not favor dimerization. Thus
dynamic preformation although it could favor dimerization, is resisted by the nearest neighbors.
The original topochemical postulate states that the
reactions in the solid state involve minimal atomic
movement. Recently, a few examples, most of which
were mentioned in section 11,have been reported which
are exceptions to the traditionally accepted criteria
regarding the alignment of reactive double bonds.
Many photochemical dimerizations, polymerization, and
excimer formations occur between double bonds or
aromatics having poor initial orbital overlap. It is unnecessary to emphasize that many of these cases could
be rationalized in the light of the dynamic
preformation concept.
N

C. Consequences of Local Stress on


Solid-State Reactions
The topochemical postulate discussed above has no
doubt been outstandingly successful, but it has its limitations. For example, &-cinnamic acid derivatives
photoisomerize in the solid state before dimerization.
More interesting exceptions which lead to an entirely
different kind of rationalization are photostimulated
reactions of azoalkanes and diacyl peroxides.27w279
There are several examples of radical pairs which yield
coupling products in solutions but give predominantly
disproportionation products in the solid state. Photolysis of azobis(3-phenyl-3-pentane)
in solution gives
3,4-diethyl-3,4-diphenylhexaneand a photostable
product (Scheme 64). However, in the crystalline state
at -78 O C disproportionation products 3-phenylpentane
and 3-phenyl-2-pentene as a 3:l mixture of the E and
2 isomers are obtained. From topochemical considerations, however, a predominance of the 2-isomer would
be expected. It follows that more than token motion
of the photolytic fragments must precede their reaction.
It was observed that azobisisobutyronitrile upon photolysis accounts for only 5 90radical-radical reactions
in the solution phase but for 95% in the crystalline
phase. From EPR studies and dynamic computer simulations it was concluded that one of the radicals
formed in a first dissociation step has to rotate about
the C-CEN axis before abstracting a hydrogen atom
from another radical (Figure 39). The difference observed in the solid-state behavior of acetyl benzoyl
peroxide from that in solution is marked. While in
solution or the melt more than a dozen products have
been identified, photolysis of crystals gave only the
simple cage products, methyl benzoate from loss of one
C02 and toluene from loss of both (Scheme 65). X-ray
studies showed that the incipient methyl radical in
acetyl benzoyl peroxide is closer to the peroxy oxygen
of the benzoyloxyl group than to its carbonyl oxygen
(3.62 and 4.07 A), but that the incipient COPmolecules
should shield the peroxy oxygen from the methyl much
more effectively than it shields the carboxyl group.

Figure 39. Schematic representation illustrating the rotational


requirement during the disproportionation of AIBN.

From an incisive study of the EPR spectra, the kinetic


anomalies of methyl phenyl and methyl benzoyloxyl
pairs were attributed to the environments mechanical
properties. When a molecule undergoes homolysis, the
molecule must undergo expansion to allow for the
cleavage of the bond and this in turn would produce
stress or pressure in the lattice. The effect of the stress
would be to influence the behavior of the radicals and
also the reactions of the surrounding molecules. The
confirmation of the stress hypothesis was obtained by
investigating the photolytic behavior of dibenzoyl
peroxide which gives a 5 K a phenyl-benzoyloxyl radical
pair. Detailed investigations of different phenomena
such as hindered rotation of neophyl radicals in 3methyl-3-phenylbutanoyl peroxide, nontopochemical
neophyl rearrangement in 3,3,3-triphenylpropanyl
peroxide and rotational translation of decyl radicals in
undecanoyl peroxide provide convincing proof in support of the hypothesis that local stress rather than topochemical factors control these reactions.279

D. Role of Defects
In a vast number of solid-state photodimerization
reactions the crystal structure of the monomer predicts
directly the stereochemistry of the dimer. But even
before topochemical control in solid-state reactions were
fully exploited in useful molecular transformations,
examples of the breakdown of such control were reported. Some of the first well-studied examples of
nontopochemically controlled reactions are the photodimerization of substituted anthracenes. In a classic
communication, Craig and Sarti-Fantoni reported280
that 9-cyanoanthracene and 9-anthraldehyde yield
head-tail photodimers whereas from the crystal structure head-head dimers would be expected. These authors suggested that in these examples reaction occurs
at defects or surfaces or in zones already disordered. It
has now become established due to the pioneering
contributions of Thomas and his co-workers that defects
play a significant role in controlling the nature of the
solid state dimerization of a n t h r a c e n e ~ . ~ ~ l - ~ ~ l
The crystal structure obtained by X-ray crystallography is an averaged structure and describes the environment of the vast majority of molecules in the solid.
This has provided the basis for topochemical arguments. However, it is known that the real crystal has
surfaces, dislocations and other defects which are not
readily detectable by X-ray diffraction methods. It is
not obvious a priori what role these minor features of
a crystal plays in controlling the nature of a photo-

Ramamurthy and Venkatesan

476 Chemical Reviews, 1987, Vol. 87, No. 2

SCHEME 65
0

reaction. Craig and Sarti-Fantoni have proposed conditions under which the influence of defects are likely
to be of importance.280 When a crystal with defects is
irradiated, radiation will initiate three events, namely,
deactivation, dimerization, and transfer of excitation
to another site. If it is assumed that the deactivation
process is independent of the nature of the site, then,
when the dimerization is slow, the process of transfer
of excitation to a neighboring site would occur with a
higher probability. Since the normal symmetry of the
sites is disrupted at the dislocation, molecules at these
sites are likely to act as trapping centers for excitation.
This means that dislocations can function as favored
areas of reaction. In other words, if the light energy can
be transferred rapidly within the crystal after absorption then the photochemistry of the ideal lattice need
not be important. Instead photoreaction would become
more probable in regions where excitation energy can
preferentially migrate. Since the number of molecules
at defect sites will be small, the reaction must be accompanied by defect multiplication to give an appreciable yield of product. Because of the importance of
transfer of energy in these systems, the reaction is very
sensitive to impurities and is readily inhibited by
suitable dopants.
The need to determine the nature of local or abnormal structure in the vicinity of structural imperfections
is, therefore, paramount. What we require is an insight
into the situation that prevails at the disrupted regions
that have been locked into the crystal either during its
genesis or as a result of subsequent deformation and
maltreatment. Further, it is necessary to establish a
correlation between sites of enhanced photoreactivity
and sites of emergence of linear defects. The extensive
tools available for attacking the problem of characterization of dislocations in metals and alloys is not suitable for the study of organic crystals. For example, in
transmission electron microscopy, by far the most
powerful technique for this purpose, it is desirable that
the specimens to be studied be chemically stable under
high-energy electron beams and relatively nonvolatile.
These requirements are not usually satisfied by organic
crystals, though the difficulties have, to some extent,
been circumvented. Topographical studies (using optical and replica electron microscopy) of cleavage and
habit faces have yielded valuable information about the
sites of emergent dislocations. Chemical etching has,
to date, proven to be the most convincing method of
locating and characterizing dislocations in organic
solids. The potential of X-ray diffraction topographyB2
in characterizing the nature of defects and their correlation with reactivity in the field of organic solid state
photochemistry has not been explored fully. An attractive feature of this method appears to be that it
would allow us to follow up the changes taking place
on the crystalline sample bathed in the X-ray beam
when it is irradiated by UV radiation. When coupled

with synchrotron sources the technique offers a promising area of activity. The value of atom-atom potential
calculations in the elucidation of photoreactions in
crystals originating in defects or in a metastable phase
already present in the parent crystalline matrix has
been recognized.289
A systematic chemical and crystallographic study in
the laboratory of Schmidt has shown that the substituted anthracenes fall into at least three packing types
In the
corresponding to a,& and y classification.293~294
y-types, the molecular planes are not parallel and the
distances between the meso atoms of neighboring
molecules are greater than 5 A. As expected these
compounds are light stable in the crystalline state
though they photodimerize in solution. In the &-type,
anthracene molecules are packed pairwise across centers
of symmetry such that the C(S).-C(lO) distances are
short (K3.6 A). 0-type packing is analogous to y except
that the meso atoms are closer (C4 A) and the molecules
pack parallel and with appreciable overlap. A few examples of anthracenes belonging to each class and their
photobehavior are summarized in Scheme 66. It is
clear from the scheme that dimerization reactions of
anthracenes generally do not conform to the topochemical principles. The photodimerization reactions
of anthracenes which do not conform to topochemical
postulates can be classified into two major categories:
(i) the monomer matrix is topochemically favorable for
dimerization, but the product is controlled by defects
or lattice imperfections, and (ii) the monomer matrix
is not suitable for topochemical dimerization, but irradiation results in dimer. In the first category lattice
imperfections could lead to products that would be
obtained either from the bulk matrix or to nontopochemical dimers (Scheme 66). In the following paragraphs a few of the systems which have been subjected
to detailed investigations are summarized.
The crystal structure of anthracene shows no molecules that are separated by C4 A, hence not appearing
to permit any reactivity. Yet photolysis of anthracene
in the crystalline state yields
Experimentally there is strong evidence to indicate that appreciable distortion of the anthracene lattice is required to
permit photodimerization. Chandross and Fergusonm7
had originally suggested that the free surface rather
than the bulk, of the crystalline anthracene functions
as the seat of the reaction. Thomas and Williams,281
supported by etch-pit studies, proposed that crystal
defects may function as the preferred centers for reaction, it being possible that anthracene molecules have
their excitation energies slightly reduced when they are
displaced from regular lattice sites. When one half of
a cleaved, melt grown crystal of anthracene is etched
to reveal the emergent dislocations and the other separated half is exposed to radiation and ultimately examined by interference contrast microscopy, the degree
of correspondence between etch pits and dimerization

Chemical Reviews, 1987, Vol. 87, No. 2 477

Photochemical Reactlons of Organic Crystals

SCHEME 66

142,L-dichlorophenoxycarbonyl)
1,s dichloro
h3_ Head-to-head dimer

9-CI
9-CHO
9-CN
1.10-dichloro

Head-to-tail dimer

7
%Lightstable

d- type
1-CI
9-CI
9-Br
9-Me
9-CO2Me
9-CONH2

-% Head-to-tail

dimer

dimerization reaction occurs at such specific sites in


crystals. Luminescence studies by Ludmer300i301
and
recently by Ebeid and Bridge302convincingly demonstrate the importance of defects in the photodimerization of 9-cyanoanthracene. Crystals of 9cyanoanthracene show an excimer emission (9-headhead configuration). Most interestingly, such emission
can be quenched in the presence of suitable dopants
anthraquinone, etc.
such as 9-chloro-10-anthraldehyde,
More importantly, dopant molecules which influence
the excimer decay of 9-cyanoanthracene also quench the
photodimerization reaction. The most important observation is the total quenching of the photoreaction
by means of the doping with 9-chloro-10-anthraldehyde.
This together with the quenching of excimer fluorescence decay proves unequivocally that the photoreaction does not occur at the site of absorption.
The crystal structure of 1,8-dichloro-lO-methylanthracene, like that of 9-cyanoanthracene, would lead
us to expect the production of head-head dimer. Yet
it is the head-tail dimer which is produced (Scheme
66).303This anomalous situation has once again been
attributed to the defects in the crystal and support for
this postulate comes from optical microscopic studies.2ss-2as It has been shown with the help of optical
microscopic investigation that dimer nuclei appear at
emergent dislocations and tend to be aligned along the
[Ool] and to a less extent, the [lo01 directions. It has
been proposed that the slip system (210) [I201 brings
about an-incipient trans configuration accounting for
the head-tail dimer formation. Crystals of 1,8-dichloro-9-methylanthracene belong to the space group
Pnma. Examination of the structure shows that the
molecules &e too far apart (- 7 A) and unfavorably
oriented to form a photodimer in the ordered crystal.
Nevertheless, irradiation gives exclusively the head-tail
dimer (Scheme 66). Efforts of Thomas and co-workers
have resulted in the identification of structural
imperfections, chiefly dislocations, in the crystals of
1,&dichloro-9-methylanthracene.It has been suggested
that slip along [OlO] is conducive for the photoproduction of the trans-dimer.
Other examples in which defects have been proposed
as the loci for photodimerization include 1,8-dichloro-,
l,kdichloro-, and 1,5-dichloroanthracenes and -acen a p h t h ~ l e n e s .For
~ ~ ~detailed descriptions of defects
in organic solids and for their importance in phototransformations readers are referred to excellent
available review^.^^"^@'

centers is very high. It is of considerable importance


to note that the recent electron microscopic study by
Thomas and co-workersm has revealed that dislocations
are probably not quite as important as was originally
thought during the photodimerization of anthracene.
Electron microscopic observations have indicated the
coexistence inside normal anthracene, of thin regions
of a metastable phase and that moderate stress produces crystallites of a new phase in coherent contact
with the parent crystal matrix in one specific direction.
It is interesting to note that in the new phase (space
group P1) the Cg-.Cg/ distance is 4.2 A whereas in the
original crystal (P2,la) this value is 4.5 A. Further,
electron microscopic observations, following UV irradiations, revealed that these crystallites of metaphase
act as nuclei for photodimerization, which, in due
course, spreads throughout the specimen. Thus, this
particular example emphasizes that the existence of
some kind of defect in organic crystals could be of little
consequence photochemically. Also it illustrates the
continued interest in seeking newer understandings for
the nontopochemical phenomenon.
9-Cyanoanthracenemolecules pack in an orthorhombic structure in which the molecules are arranged in
columns with an interplanar distance of 3.5 A. The
marked overlap between molecules in the stack makes
this structure ideal for head-head dimer formation.
V II . Conclusion
However, irradiation of 9-cyanoanthracene crystals results in a photodimer with head-tail geometry. This
Interest in solid-state reactivity is not as widespread
as solution-phase studies and remains in the cradle of
so called violation of the topochemical principle has
a limited number of investigators. Although the main
been interpreted as being due to a photoreaction at
factors required for the dimerization reaction to occur
defect sites.280Interference contrast and fluorescence
in the solid state have been very well understood and
microscopy have been employed by Cohen, Thomas,
the subtleties of the topochemical factors have come to
and their co-workers for the examination of cleaved and
partially dimerized faces of the m o n ~ m e r . As
~ ~in~ , ~ ~light from recently reported results, the strategies for
aligning the reactive bonds still remain to be achieved.
the case of anthracene, discussed previously, a good 1:l
The future strength of the field will be clearly detercorrespondence between regions of etch pita and regions
of reaction was also observed in 9-cyanoanthracene
mined by increased progress in our understanding of
crystals. Further, it was suggested that along the (221)
the factors controlling molecular packing. On the exslip plane preformation of monomer molecules having
perimental side, apart from looking for chemical groups
the antiregistry is feasible and that the photowhich may be of potential value in steering molecules

Ramamurthy and Venkatesan

478 Chemical Reviews, 1987, Vol. 87, No. 2

in the crystal lattice, other avenues for achieving the


topochemical requirements need to be explored. In this
connection the strategies which are being employed for
achieving segregated stacks of molecules in organic
metals may be worth trying. Further, the possibility
of using special topological features of molecules such
as organic and inorganic systems as hosts exists and
may prove fruitful.
It is also clear from the present review that although
some successful attempts with regard to asymmetric
synthesis in the solid state have been reported, further
development in this very important area has not progressed and demands attention. There is a clear lack of
understanding the mechanism of solid-state reactions.
Present knowledge exists only on the initial and final
stages of the organic solid-state reactions. Progress in
this direction is expected in the coming years. Finally,
many new solid-state reactions need to be discovered
and many remarkable reactions reported in the beginning of the century should be investigated thoroughly.
It seems likely that properly understanding reactions
in organic crystals and fully realizing their synthetic
potential will require better insight into the chemical
and physical properties of solids at the molecular level.
It is hoped that the knowledge gained through the
studies of organic reactions in crystals, in turn, may be
relevant to the solution of a number of pending problems such as the development of organic materials
possessing unusual electrical and optical properties.

Acknowledgments. Our thanks are due to Drs. P.


Arjunan and G. Usha and G. S. Murthy, K. Padmanabhan, and D. Kanagapuspam for their help in the
preparation of the manuscript. We greatly appreciate
Prof. J. R. Scheffer for constructive suggestions and for
continued encouragement. The University Grants
Commission, Government of India is thanked for financially supporting our solid-state program.
References
Stobbe, H.; Steinberger, F. K. Chem. Ber. 1922, 55, 2225.
Ribber, Chem. Ber. 1902, 35, 2411.
Ciamician, G.; Silber, P. Chem. Ber. 1902, 35, 4128.
Ramamurthy, V. Tetrahedron 1986, 42, 5753.
Haseeawa. M. Chem. Rev. 1983.83.507.
Dill&.
..
.
.
.
. W. L. Chem. Reu. 1983.83. 1.
HaseGwa, M.Adv. Poly. Sci. 1982; 42, 1.
Schmidt, G. M. J. Pure Appl. Chem. 1971,27, 647.
Cohen, M. D.; Green, B. S. Chem. Br. 1973,9, 490.
Thomas. J. M. Philos. Trans. R. SOC.
London 1974,277,251.
Thomas: J. M.: Morsi. S.E.: Desverene. J. P. Adv. Phv.
- Ore.
Chem. 1977,15, 63.
Thomas, J. M. Pure Appl. Chem. 1979,51, 1065.
Addadi, L.; Ariel, S.; Lahav, M.; Leiserowitz, L.; PopovitzBiro, R.; Tang, C. P. Chemical Physics of Solids and their
Surfaces; Specialist Periodical Reports; Royal Society: London, 1979; Vol. 8, p 202.
Schmidt, G. M. J. et. al. Solid State Photochemistry Ginsburg, D., Ed.; Verlag Chemie: New York, 1976.
Cohen, M. D.; Schmidt, G. M. J. J. Chem. SOC.
1964, 1996.
Cohen, M. D.; Schmidt, G. M. J.;Sonntag, F. I. J . Chem. SOC.
1964, 2000.
Schmidt, G. M. J. J. Chem. SOC.
1964, 2014.
Kohlshutter, H. W. 2. Anorg. Allg. Chem. 1918, 105, 121.
Gnanaguru, K.; Ramasubbu, N.; Venkatesan, K.; Ramamurthy, V. J. Org. Chem. 1985,50, 2337.
Ramasubbu, N.; Gnanaguru, K.; Venkatesan, K.; Ramamurthy, V. Can. J. Chem. 1982,60, 2159.
Bhadbhade, M. M.; Murthy, G. S.; Venkatesan, K.; Ramamurthy, V. Chem. Phys. Lett. 1984, 109, 259.
Gnanaguru, K.; Murthy, G. S.;Venkatesan, K.; Ramamurthy,
V. Chem. Phys. Lett. 1984,109, 255.
Gnanaguru, K.; Ramasubbu, N.; Venkatesan, K.; Ramamurthy, V. J . Photochem. 1984,27, 355.

(24) Ramasubbu, N.; Guru Row, T. N.; Venkatesan, K.; RamaChem. Commun.
murthy, V. Rao, C. N. R. J. Chem. SOC.,
1982, 178.
(25) Nakanishi, H.; Jones, W.; Thomas, J. M. Chem. Phys. Lett.
1980, 71, 44.
(26) Nakanishi, H.; Jones, W.; Thomas, J. M.; Hursthouse, M. B.;
Chem. Commun. 1980, 611.
Motevalli, M. J. Chem. SOC.,
(27) Jones, W.; Nakanishi, H.; Theocharis, C. R.; Thomas, J. M.
J. Chem. SOC.,
Chem. Commun. 1980, 610.
(28) Thomas. J. M. Nature (London) 1981.289. 633.
(29) Nakanishi, H.; Jones, W:; Thomas, J. hi.;Hksthouse, M. B.;
Motevalli, M. J. Phys. Chem. 1981, 85, 3636.
(30) Jones, W.; Ramdas, S.; Theocharis, C. R.; Thomas, J. M.;
Thomas, N. W. J . Phys. Chem. 1981,85, 2594.
(31) Sadeh, T.; Schmidt, G. M. J. J . Am. Chem. SOC.
1962, 84,
3970.
B 1967, 239.
(32) Lahav, M.; Schmidt, G. M. J. J . Chem. SOC.
B 1967, 312.
(33) Lahav, M.; Schmidt, G. M. J. J. Chem. SOC.
(34) Green, B. S.; Lahav, M.; Schmidt, G. M. J. J . Chem. SOC.
E
1971, 1552.
(35) Pfoertner, K. H.; Englert, G.; Schoenholzer, P. Abstracts of
Papers, 12th International Conference on Photochemistry,
Tokyo, Japan, 1985; p 487. Pfoertner, K. H.; Englert, G.;
Schoenholzer. P. Tetrahedron. in Dress.
(36) Nakanishi, F.; Nakanishi, H.; Tsuchiya, M.; Hasegawa, M.
Bull. Chem. SOC.
Jpn. 1976,49, 3096.
(37) Nakanishi, H.; Hasegawa, M.; Mori, T. Acta Crystallogr.
Sect. C: Cryst. Struit. Commun. 1985, C41, 70.
(38) Nakanishi, H.; Parkinson, G. M.; Jones, W.; Thomas, J. M.;
Hasegawa, M. Isr. J. Chem. 1979, 18, 261.
(39) Ariel, S.;
Askari, S.; Scheffer, J. R.; Trotter, J.; Walsh, L. J.
Am. Chem. SOC.
1984, 106, 5726.
(40) Hanson, A. W. Acta Crystallogr.Sect. B Struct. Crystallogr.
Cryst. Chem. 1975, B31, 1963.
(41) Sen, N.; Venkatesan, K., unpublished results.
(42) Mez, H.; Rihs, G. Helv. Chim. Acta 1973, 56, 2766.
(43) Sinnreich, J.; Batzer, H. Helv. Chim. Acta 1973, 56, 2760.
(44) Kanagapushpam, D.; Venkatesan, K.; Ramamurthy, V. Acta
Cryst;aiiogr.; in press.
Theocharis, C. R.; Jones, W.; Thomas, J. M.; Motevalli, M.;
Hursthouse, M. B. J. Chem. SOC.,Perkin Trans. 2 1984,71.
Kearsley, S.K.; Desiraju, G. R. Proc. R. SOC.
London, Ser. A
1985, 397, 157.
(a) Frank, J. K.; Paul, I. C. J. Am. Chem. SOC.
1973,95,2324.
(b) Leonard, N. J.; McCredie, R. S.;
Logue, M. W.; Cundall,
R. L. Ibid. 1973. 95. 2320.
(48) Irngartinger, H.; Aeker, R. D.; Rebafka, W.; Staab, H. A.
Angew. Chem., Int. Ed. Engl. 1974,13,674.
(49) Hasegawa, M.; Nohara, M.; Saigo, K.; Mori, T.; Nakanishi,
H. Tetrahedron Lett. 1984,25,561. Hasegawa, M.; Saigo, K.;
Mori. T.: Uno. H.: Nohara.. M.:. Nakanishi, H. J. Am. Chem.
SOC.
1985, 107, 2788.
(50) Kaftory, M. J. Chem. SOC.,Perkin Trans. 2 1984, 757.
(51) Patel, G. N.; Duesler, E. N.; Curtin, D. Y.; Paul, I. C. J . Am.
Chem. SOC.
1980, 102, 461.
(52) Wegner, G. Pure Appl. Chem. 1977, 49, 443. Misin, V.;
Cherkashin. RUSS.Chem. Rev. 1985,54,562.
(53) Blasky, V. K.; Zorkii, P. M. Acta Crystallogr. Sect. A Cryst.
Phys., Diff., Theor. Gen. Crystallogr. 1977,33A, 1004.
(54) Addadi, L.; Cohen, M. D.; Lahav, M. J. Chem. SOC.,Chem.
Commun. 1975,471.
(55) Addadi, L.: Cohen, M. D.; Lahav, M. Mol. Cryst. Liq. Cryst.
1976, 32, 137.
(56) Addadi, L.; Lahav, M. J. Am. Chem. SOC.
1978, 100, 2838.
(57) Addadi, L.; Gati, E.; Lahav, M.; Leiserowitz, L. Zsr. J. Chem.
1976-77,15, 116.
(58) Addadi, L.; Lahav, M. J. Am. Chem. SOC.1979, 101, 2152.
(59) Addadi, L.; van Mil, J.; Lahav, M. J . Am. Chem. SOC.1982,
104,3422.
(60) van Mil. J.: Addadi. L.: Lahav. M.: Leiserovitz. L. J . Chem.
SOC.,Chem. Commun.1982, 584.
(61) Elgavi, A.; Green, B. S.;
Schmidt, G. M. J. J. Am. Chem. SOC.
1973, 95, 2058.
Lahav, M.; Schmidt, G. M. J. Mol. Cryst. Liq.
(62) Green, B. S.;
Cryst. 1975,29, 187.
(63) Green, B. S.; Lahav, M.; Rabinovich, D. Acc. Chem. Res.
1979, 12, 191.
(64) Addadi, L.; Lahav, M. Pure Appl. Chem. 1979, 51, 1269.
(65)
Cryst.
. . Holland, H. L.; Richardson, M. F. Mol. Cryst. Lis.
.~
1980,58; 311.
(66) Chenchaiah, P. C.; Holland, H. L.; Richardson, M. F. J .
Chem. SOC.,Chem. Commun. 1982, 436.
(67) Rabinovich, D.; Schmidt, G. M. J. J. Chem. SOC.
B 1967, 144.
1961,
(68) Cookson, R. C.; Cox, D. A.; Hudec, J. J. Chem. SOC.
4499.
(69) Kaupp, G.; Jostkleigrewe, E.; Hermann, H. J. Angew. Chem.,
Int. Ed. Engl. 1982, 21, 435.
(70) Chase, D. B.; Amey, R. L.; Holtje, W. G. Appl. Spectrosc.
1982, 36, 155.
(71) Waschen, E.; Matusch, R.; Krampity, D.; Hartke, K. Liebigs
Ann. Chem. 1978, 2137.
~I

Chemical Reviews, 1987, Vol. 87, No. 2 479

Photochemical Reactions of Organic Crystals

Griffin, G. W.; Basinski, J. E.; Vellturo, A. F. Tetrahedron


Lett. 1960, 13.
Griffin, G. W.; Vellturto, A. F.; Furukawa, K. J. Am. Chem.
SOC.1961,83, 2725.
Williams, J. L. R. J. Org. Chem. 1960, 25, 1839.
Miller. D. B.: Flanaean. P. W.: Scechter. H. J. Am. Chem.
SOC.1972, 94, 3912.Donati, D.; Fiorenza, M.; Sarti-Fantoni, P. J. J. Heterocycl.
Chem. 1979,16, 253.
Donati, D.; Fiorenza, M.; Moschi, E.; Sarti-Fantoni, P. J. J.
Heterocycl. Chem. 1977, 14, 951.
Ried, W.; Bopp, H. Angew. Chem., Int. Ed. Engl. 1977, 16,
653.
Taylor, E. C.; Paudler, W. W. Tetrahedron Lett. 1960,25, 1.
Wang, S. Y. Nature (London) 1963,200, 879.
Powell, B. M.; Martel, P. Photochem. Photobiol. 1977, 26,
305.
DeBoer, C. D.; Schlessinger, R. H. J. Am. Chem. SOC.1968,
90. 803.
Bremner, J. B.; Warrener, R. N.; Adman, E.; Jensen, L. H. J.
Am. Chem. SOC.1971,93,4574.
Bates, R. B.; Christensen, K. A,; Hallberg, A.; Klenck, R. E.;
Martin, A. R. J. Org. Chem. 1984, 49, 2978.
Bedev. M. J.: Crombie. L.: KnaDD.
.. T. F. W. B. J.Chem. SOC..
PeFkLi Trans. 1 1979.'976.
Mohr, S. Tetrahedron Lett. 1979, 3139.
Adam, G. Tetrahedron 1973,29, 3177.
Kutschabsky, L.; Reck, G.; Adam, G. Tetrahedron 1975,31,
'

3065.

Adam, G. Tetrahedron Lett. 1971, 1357.

Adam, G.; Voigt, B. Tetrahedron Lett. 1971, 4601.


Cookson, R. C.; Crundwell, E.; Hill, R. R.; Hudic, J. J. Chem.
SOC.1964, 3062.
Cookson, R. C.; Hill, R. R.; Hudec, J. J. Chem. SOC.1964,
3043.

Houk, K. N.; Northington, D. J. Tetrahedron Lett. 1972,303.


Lahav, M.; Schmidt, G. M. J. Tetrahedron Lett. 1966,2957.
Uhler, R. 0.; Tiers, G. V. D. J. Am. Chem. SOC.1962, 84,
3397.
Berkovitch-Yellin, Z.; Lahav, M.; Leiserovitz, L. J. Am.
Chem. SOC.
1974, 96,918.
Perlmutter, H. D.; Trattner, R. B. J. Org. Chem. 1978, 43,
3n.5~

kielke, R. D.; Copenhafer, R. A. Tetrahedron Lett. 1971,879.


Nishio, T.; Nakajima, N.; Omote, Y. Tetrahedron Lett. 1980,
21, 2529.
Hazell, A. C.; Pagni, R. M.; Persy, G.; Rommel, E.; Wirz, J.
Helv. Chim. Acta 1981, 64, 2830.
Fuchs, E.; Pasternak, M. J. Chem. SOC.,Chem. Commun.
1977, 537.
Mohr, S. Tetrahedron Lett. 1979, 2461.
Lawrenz, D.; Mohr, S.; Wendlander, B. J.Chem. SOC.,Chem.
Commun. 1984, 863.
Bradshaw, C. K.; Bearers, L. E.; Jones, J. H. J. Org. Chem.
1957,22, 1740.
Vansant, J.; Toppet, S.; Smets, G.; Declercq, J. P.; Germain,
G.; van Meerssehe, M. J. Org. Chem. 1980, 45, 1565.
Marciani, S.; Dallacqua, F.; Rodighiero, P.; Caporalle, G.;
Rodighiero, G. Gazz. Chim. Ital. 1970, 100,435.
Rodighiero, G.; Cappellina, V. Gazz. Chim. Ital. 1961,91,103.
Quina, F. H.; Whitten, D. G. J. Am. Chem. SOC.1975, 97,
1602.
Guru Row, T. N.; Ramachandra Swamy, H.; Ravi Acharya,
K.; Ramamurthy, V.; Venkatesan, K.; Rao, C. N. R. Tetrahedron Lett. 1983,24,3263. Ramachandra Swamy, H.; Guru
Row, T. N.; Ramamurthy, V.; Venkatesan, K.; Rao, C. N. R.
Curr. Sci. 1982,51, 381. Ramachandra Swamy, H.; Ramamurthy, V.; Venkatesan, K. Indian J. Chem. 1982,2IB, 79.
Werbin, H.; Strom, E. T. J. Am. Chem. SOC.1968,90,7296.
Nalini, V.; Desiraju, G. R. Tetrahedron, in press. Kato, S.;
Nakatani, M.; Harashina, H.; Saigo, K.; Hasegawa, M.; Sato,
S. Chem. Lett. 1986, 847. Hasegawa, M.; Arioka, H.; Harashina, H.; Nohara, M.; Kubo, M.; Nishikubo, T. Isr. J. Chem.
1985. 25. 302.
Nakanishi, F.; Nakanishi, H.; Suzuki, Y.; Hasegawa, M.
Chem. Lett. 1974, 525.
Nakanishi, F.; Yamada, S.; Nakanishi, H. J. Chem. SOC.,
Chem. Commun. 1977, 247.
Nakanishi, F.: Hirakawa, H.: Nakanishi, H. Isr. J . Chem.
1979, 18, 295.
Scheffer, J. R. Acc. Chem. Res. 1980, 13, 283. Shklover, V.
E.; Timofeeva, T. V. Russ. Chem. Rev. 1985,54,619.
Dzakpasu, A. A.; Phillips, S. E. V.; Scheffer, J. R.; Trotter,
J. J . Am. Chem. SOC.1976, 98, 6049.
Scheffer, J. R.; Dzakpasu, A. A. J. Am. Chem. SOC.1979,100,
2163.
Weisz, A.; Kaftory, M.; Vidavsky, I.; Mandelbaum, A. J.
Chem. SOC.,Chem. Commun. 1984, 18.
Appel, W. K.; Greenhough, T. J.; Scheffer, J. R.; Trotter, J.;
Walsh, L. J. Am. Chem. SOC.1980, 102, 1158.

(120) Appel, W. K.; Greenhough, T. J.; Scheffer, J. R.; Trotter, J.;


Walsh, L. J. Am. Chem. SOC.1980,102, 1160.
(121) Appel, W. K.; Jiang, Z. Q.; Scheffer, J. R.; Walsh, L. J.Am.
Chem. SOC.1983,105, 5354.
(122) Jiang, Z. Q.; Scheffer, J. R.; Secco, A. S.; Trotter, J. Tetrahedron Lett. 1981,22, 891.
(123) Greenhough, T. J.; Scheffer, J. R.; Secco, A. S.; Trotter, J.;
Walsh. L. Isr. J. Chem.. in Dress.
(124) Ariel, S.; Evans, S.; Hwang,
Jay, J.; Scheffer, J. R.; Trotter, J.; Wong, Y. F. Tetrahedron Lett. 1985, 26, 965.
(125) Appel, W. K.; Greenhough, T. J.; Scheffer, J. R.; Trotter, J.
J.-Am. Chem. SOC.1979; 101, 213.
(126) Matsuura, T.; Sata, Y.; Ogura, K. Tetrahedron Lett. 1968,
4627.
(127) Williams, J. R.; Abdel-Magid, A. Tetrahedron 1981,37,1675.
(128) Sobti, R. R.; Levine, S. G.; Bordner, J. Acta Crystallogr. Sect.
B: Struct. Crystallogr. Cryst. Chem. 1972, B28, 2292.
(129) Mc Phail, A. T.; Luhan, P. A.; Ts-Chang, P. S. W.; Onan, K.
D. J. Chem. SOC.,Perkin Trans. 2 1977, 379.
(130) Aoyama, H.; Hasegawa, T.; Omote, Y. J. Am. Chem. SOC.

e.;

1979. 101. 5343.


7

- - - I

(131) Ariel, S.; Ramamurthy, V.; Scheffer, J. R.; Trotter, J. J. Am.


Chem. SOC.1983,105, 6959.
(132) Wagner, P. J.; Giri, B. P.; Scaiano, J. C.; Ward, D. L.; Gabe,
E.: Lee. F. L. J. Am. Chem. SOC.1985.107. 5483.
(133) Mohr, S. Tetrahedron Lett. 1980,21,593. '
(134) Mohr, S. Tetrahedron Lett. 1979, 3139.
(135) Wagner, P. J. Top. Curr. Chem. 1976, 66, 1.
(136) Evans, S.; Omkaram, N.; Scheffer, J. R.; Trotter, J. Tetrahedron Lett. 1985,26, 5903.
(137) Gore, P. H.; Hoskins, J. A.; Lott, K. A. K.; Waters, D. N.
Photochem. Photobiol. 1970, 11, 551. Migirdicyan, E.;
Despres, A.; Lejeune, V.; Leach, S. J. Photochem. 1974-75,
3, 383.
(138) Ito, Y.; Matsuura, T. Tetrahedron, in press.
(139) Bamberger, E.; Elger, F. Ann. Chem. 1909, 371, 319.
(140) Tanasescu, I.; Tanasescu, H. Chem. Abstr. 1926, 20, 749.
(141) Steiger, R. Helv. Chim. Acta 1933, 16, 793.
(142) Schultz, G.; Ganguly, K. L. Ber. Dtsch. Chem. Ges. 1925,58,
702.
(143) Berson, J. A.; Brown, E. J. Am. Chem. SOC.1955, 77, 447.
(144) Tanasescu, I. Chem. Abstr. 1925, 19, 2932.
(145) Tanasescu, I.; Macovsky, E. Chem. Abstr. 1930, 24, 2429.
(146) Ried, W.; Wilk, M. Ann. Chem. 1954,590, 91.
(147) Sachs, F.; Hilpert, S. Ber. Dtsch. Chem. Ges. 1905,37,3425.
(148) Sachs, F.; Hilpert, S. Ber. Dtsch. Chem. Ges. 1906, 39, 899.
(149) Krohnke, F.; Krohnke, G.; Vogt, I. Chem. Ber. 1953,86,1500.
(150) Trotter, J. Acta Crystallogr. 1959, 12, 605.
(151) Carper, W. R.; Davis, L. P.; Extine, M. W. J. Phy. Chem.
1982. 86.
- ., 459.
.
-- .
(152) Dopp, D. Org. Photochem. Synth. 1976, 2, 43.
(153) Douu, D.; Sailer, K.-H. Chem. Ber. 1975, 108, 3483.
(154) Dopp, D. Tetrahedron Lett. 1971, 2757.
(155) Dopp, D.; Sailer, K.-H. Tetrahedron Lett. 1971, 2761.
(156) Dopp, D.; Sailer, K.-H. Chem. Ber. 1975, 108, 301.
(157) Padmanabhan, K.; Dopp, D.; Venkatesan, K.; Ramamurthy,
V. J. Chem. SOC.,
Perkin Tram. 2 1986,897. Padmanabhan,
K.; Ramamurthy, V.; Venkatesan, K.; Ponnuswamy, M. N.;
Trotter, J. Acta Cryst. Sect. C. Cryst. Struct. Commun. 1986,
42, 610.
(158) Padmanabhan, K.; Schmidt, R.; Dopp, D.; Venkatesan, K.;
Ramamurthy, V. J. Chem. SOC.,Perkin Trans. 2, in press.
(159) Senier, A.; Shepheard, F. G. J. Chem. SOC.1909, 95, 1943.
(160) Senier, A.; Shepheard, F. G.; Clarke, R. J . Chem. SOC.1912,
101, 1950.
(161) Senier, A.; Shepheard, F. G. J. Chem. SOC.1909, 95, 441.
(162) Senier, A.; Forster, R. B. J. Chem. SOC.1914, 105, 2462.
(163) de Gaouck, V.; le Fevre, R. J. W. J. Chem. SOC.1939, 1457.
(164) Schmidt, G. M. J. Reactivity of the Photoercited Organic
Molecules; Interscience: New York, 1967; pp 227-288.
(165) Cohen, M. D.; Schmidt, G. M. J. In Reactivity of Solids; de
Boer., Ed.; Elsevier: Amsterdam, 1961; pp 556.
(166) Cohen, M. D.; Schmidt, G. M. J.; Flavian, S. J. Chem. SOC.
1964, 2041.
(167) Cohen, M. D.; Hinhberg, Y.; Schmidt, G. M. J. J.Chem. SOC.
1964, 2051.
(168) Cohen, M. D.; Hinhberg, Y.; Schmidt, G. M. J. J.Chem. SOC.
1964, 2060.
(169) Bregman, J.; Leiserowitz, L.; Schmidt, G. M. J. J . Chem. SOC.
1964, 2068.
(170) Bregman, J.; Leiserowitz, L.; Osaki, K. J. Chem. SOC.1964,
'

2086.

(171) Higelin, D.; Sixi, H. Chem. Phys. 1983, 77, 391.


(172) Hadjoudis, E.; Moustakali-Mavridis, I.: Xexakis, J. Isr. J.
Cheh. 1979.18, 202.
(173) Bernstein, J.: Schmicit. G. M. J. J. Chem. SOC..Perkin Tram.
2 1972, 951.
(174) Quinkert, G.; Tabata, T.; Hickmann, E. A. J.; Dobrat, W.
Angew. Chem., Int. Ed. Engl. 1971, 10, 199.
(175) Quinkert, Q.; Opitz, K.; Wiersdorff, W. W.; Weinlich, J.
'

4)

Shemical Reviews, 1987, Vol. 87, No. 2

Tetrahedron Lett. 1963, 1863.

( 1 7 - , Baretz, B. H.;Tuno, N. J. J . Am. Chem. SOC.1983,105,1309.


(17 Tomioka, H.; Izawa, Y. J . Chc 1 . SOC.,Chem. Commun. 1980,
I

(178)
(179)
(180)
(181)
(182)
(183)
(184)
(185)
(186)
(187)
(188)
(189)
(190)
(191)
(192)
(193)
(194)
(195)
(196)

446.
Hoppe, W.; Rauch, R. 2. Kr ,allogr. 1960, 115, 141.
Weigert, F. 2. Elektrochem. 318,24, 222.
Marckwald. W. 2. Phvs. Chc 1. 1899. 30. 140.
Scheibe, G.; Feichtmiyr, F. J. Phys. CAem. 1962,66, 2442.
Gutowsky, H. S.;
Rutledge, R. L.; Hunsberger, J. H. J . Chem.
Phys. 1958,29, 1183.
Hub, W.; Shneider, S.; Dorr, F. Angeu.Chem., Znt. Ed. Engl.
1979, 18, 323.
Kaupp, G.; Zimmerman, I. Angew. Chem., Znt. Ed. Engl.
1981,20, 1018.
Stobbe, H. Justus Liebigs Ann. Chem. 1911, 380, 1.
Hanel, H. Naturwissenchaften 1950, 37, 91.
Stobbe, H. Justus Liebigs Ann. Chem. 1908, 359, 1.
Santiago, A.; Becker, R. S.J. Am. Chem. SOC.1968,90,3654.
Ichimura, K.; Watanabe, S.Bull. Chem. SOC.Jpn. 1976,49,
2220.
Ichimura, K.; Watanabe, S. Tetrahedron Lett. 1972, 821.
Dopp, D.; Muller, D. Tetrahedron Lett. 1978, 3863.
Canon, J. R.; Patrick, V. A.; Raston, C. L.; White, A. H. A u t .
J . Chem. 1978, 31, 1265.
Ciamician, G.; Silber, P. Ber. Dtsch. Chem. Ges. 1901, 34,
2040.
Cohn, P.; Friedlander, P. Ber. Dtsch. Chem. Ges. 1902, 35,
1265.
Pleiffer, P. Ber. Dtsch. Chem. Ges. 1912, 45, 1819.
Pleiffer, P.; Kramer, E. Ber. Dtsch. Chem. Ges. 1913, 46,
3665

(197) Liighton, P. A,; Lucy, F. A. J . Chem. Phys. 1934, 2, 756.


(198) Lucy, F. A.; Leighton, P. A. J. Chem. Phys. 1934, 2, 760.
(199) Brown, E. J.; Hartley, H.; Scott, W. D.; Watts, H. G. J. Chem.
SOC.1924. 125. 1218.
Ruggli, P.JustusLiebigs Ann. Chem. 1912, 392, 92
Sachs, F.; Hilpert, S.Ber. Dtsch. Chem. Ges. 1904,3; 3425.
Senier, A.; Clarke, R. J. Chem. SOC.1914, 105, 1917.
Pleiffer, P. Justus Liebigs Ann. Chem. 1954, 590, 91, 111.
de Mayo, P.; Reid, S. T. Q.Reu. 1961, 15, 393.
Coppens, P. Acta Crystallogr. 1964, 17, 573.
Coppens, P.; Schmidt, G. M. J. Acta Crystallogr. 1964, 17,
222.
Engler, C.; Dorant, K. Chem. Ber. 1895,28, 2497.
Jungk, A. E.; Luwish, M.; Pinchas, S.; Schmidt, G. M. J. Zsr.
J. Chem. 1977,16, 308.
Ellam, R. M.; East, P. B.; Kelly, A.; Khan, R. M.; Lee, J. B.;
Linsey, D. C. Chem. Znd. 1974, 74.
Badger, G. M.; Buttery, R. G. J. Chem. SOC.1954, 2243.
Porter, C. W.; Wilbur, P. J . Am. Chem. SOC.1927,49, 2145.
Comer, F. W.; Trotter, J. J. Chem. SOC.B 1966, 11.
Lonsdale, D. K.; Nare, E.; Stephens, J. F. Proc. R. SOC.London, A 1966,261, 1.
Freer, A. A.; Mac Alpine, D. K.; Peacock, J. A.; Porte, A. L.
J. Chem. SOC.,Perkin Trans. 2 1985, 971.
Bregman, J.; Osaki, K.; Schmidt, G. M. J.; Sonntag, F. I. J .
Chem. SOC.1964, 2021.
Pryan, R. F.; Hartley, P. J. Chem. SOC.,Perkin Trans. 2 i?L,
il.

riffin, G. W.; OConnel, E. J.; Kelliher, J. M. Proc. Chc , i .


SOC.1964, 337.
Bart. J. C. J.: Schmidt. G. M. J. Rec. Trau. Chim. Pavs-Bas
1978, 97, 231.
Gougoutas, J. Z.; Naae, D. G. J . Phys. Chem. 1978,82,393.
Gougoutas, J. Z. J . Am. Chem. SOC.1979,101, 5672. Gougoutas, J. Z.; Toeplitz, B. K. Cryst. Struct. Commun. 1977,6,
331.
Evans, S. V.; Garcia-Garibay, M.; Omkaram, N.; Scheffer, J.
R.; Trotter, J.; Wireko, F. J. Am. Chem. SOC.1986,108,5648.
Paul, I. C.; Curtin, D. Y. Acc. Chem. Res. 1973, 6, 217.
Paul, I. C.; Curtin, D. Y. Science, (Washington, D.C., I883-)
1975, 187, 19.
Hochstrasser, R. M. Can. J. Chem. 1959, 37, 1123.
Brenner, G.; Roberts, F. E.; Hoinowski, A.; Budavari, J.;
Powell, B.; Hinkley, D.; Schoenewaldt, E. Angew. Chem., Znt.
Ed. Engl. 1969, 8 , 975.
Lewbart, M. L. Nature (London) 1969,222, 663.
Lin, C. T.; Perrier, P.; Clay, G. G.; Sutton, P. A,; Byrn, S. R.
J . Org. Chem. 1982, 47, 2978.
Arjunan, P.; Ramamurthy, V.; Venkatesan, K. J. Org. Chem.
1984,49, 1765.
Desvergne, J. P.; Bouas-Laurent, H.; Blackburn, E. V.; Lapouyade, R. Tetrahedron Lett. 1974, 947.
Perkin Trans.
Desvergne, J. P.; Thomas, J. M. J . Chem. SOC.,
2 1975, 584.
Shoppee, C. W. J . Chem. SOC.,Perkin Trans. 1 1985, 45.
Cohen, M. D.; Green, B. S.; Ludmer, Z.; Schmidt, G. M. J.
Chem. Phys. Lett. 1970, 7, 486.
Greer, B. S.; Schmidt, G. M. J. Tetrahedron Lett. 1970,4249.
Cohen, M. D.; Elgavi, A.; Green, B. S.; Ludmer, Z.: Schmidt,
G. M. J. J . Am. chem. SOC.1972, 94, 6776.
I

Ramamurthy and Venkatesan

(235) Hung, J. D.; Lahav, M.; Luwisch, M.; Schmidt, G. M. J. Isr.


J. Chem. 1972,10, 585.
(236) Cohen, M. D.; Cohen, R.; Lahav, M.; Nie, P. L. J. Chem. SOC.,
Perkin Trans. 2 1973, 1095. Green, B. S.; Heller, L. J . Org.
Chem. 1974.39. 196.
(237) Desiraju, G.R.;Sarma, J. A. R. P. Acc. Chem. Res. 1986,19,
222. Ramasubbu, N.; Parthasarathy, R.; Murray-Rust, P. J .
Am. Chem. SOC.1986,108, 4308.
(238) Thomas, N. W.; Desiraju, G. R. Chem. Phys. Lett. 1984,110,
99.
(239) Sarma, J. A. R. P.; Desiraju, G. R. Chem. Phys. Lett. 1985,
117, 160.
(240) Kitaigorodsky, A. I. Molecular Crystals and Molecules; Academic Press: New York, 1973.
(241) Murthy, G. S.; Venkatesan, K.; Ramamurthy, V., unpublished results.
(242) Nakanishi, H.; Sasada, Y. Acta Crystallogr. Sect. B: Struct.
Crystallogr. Cryst. Chem. 1978, B34, 332.
(243) Ueno, K.; Nakanishi, H.: Haseeawa. M.: Sasada. Y. Acta
Crystallogr. Sect. B Struct. Cry&allogr. Cryst. Chem. 1978,
B34. 2034.
(244) Desiraju, G. R.; Kamala, R.; Hanuma Kumari, B.; Sarma, J.
A. R. P. J. Chem. SOC.,Perkin Trans. 2 1984, 181.
1 Lewis, F. D.; Oxman, J. D.: Huffman, J. C. J. Am. Chem. SOC.
1984,106,466.
Alcock, N. W.; de Meester, P.; Kemp, T. J. J. !::em. SOC.,
Perkin Trans. 2 1979, 921.
Praetorius, P.; Kohn, F. Ber. Dtsch. Chem. G, 1910, 43,
2744.
Theocharis, C. R.; Desiraju, G. R.; Jones, W. J. Am. Chem.
SOC.1984,106,3606. Jones, W.; Theocharis, C. R.; Thomas,
J. M.; Desiraju, C. R. J . Chem. SOC.,Chem. Commun. 1983,
1443.
Sarma, J. A. R. P.; Desiraju, G. R. J . Am. Chem. SOC.
1986,
108, 2791. Sarma, J. A. R. P.; Desiraju, G. R. J . Chem. SOC.,
Perkin Trans. 2 1985, 1905.
Thomas, N. W.; Ramdas, S.;Thomas, J. M. Proc. R. SOC.
London, A 1985, 400, 219.
Cohen, M. D. Angew. Chem., Znt. Ed. Engl. 1975, 14, 386.
Cohen, M. D. Mol. Cryst. Liq. Cryst. 1979,50, 1.
Collins, M. A.; Craig, D. P. Chem. Phys. 1981, 54, 305.
Craig, D. P.; Mallet, C. P. Chem. Phys. 1982, 65, 129.
Craig, D. P.; Lindsay, R. N.; Mallet, C. P. Chem. Phys. 1984,
89. 187.

Warschel, A.; Huller, E. Chem. Phys. 1974, 6, 465.


Cohen, M. D.; Haberlcorn, R.; Huller, E.; Ludmer, Z.; Michel-Beyerle, M. E.; Rabinovich, D.; Sharon, R.; Warshel, A.;
Yakhot, V. Chem. Phys. 1978,27, 211.
Warshel, A.; Shakked, Z. J. Am. Chem. SOC.1975,97,5679.
Cohen, M. D.; Klein, E.; Ludmer, Z. Yakhot, V. Chem. Phys.
1974. 5. 15.
Cohen,M. D.; Yakhot, V. Chem. 1- ~ y s 1974,5,
.
27.
Cohen, M. D.; Yakhot, V. Chem. Phys. 1974,5,478.
Ariel, S.;Askari, S.: Scheffer, J. R.: Trotter. J.: Walsh. L. In
Organic Phototransformations in Nonhomogeneous Uedia;
Fox, M. A,, Ed.; ACS Symposium Series 278; American
Chemical Society: Washington, DC, 1985; pp 243-256.
Arjunan, P.; Murthy, G. S.; Venkatesan, K.; Ramamurthy, V.
Tetrahedron, in press.
Busing, W. R. WIMIN, a Computing program to Model Molecules and Crystals in Terms of Potential Energy Function,
Oak Ridge National Laboratory: Oak Ridge, 1981.
Gavezzotti, A. J.Am. Chem. SOC.
1983,105,5220. Gavezzotti,
A.; Simonetta, M. Chem. Reu. 1982,82, 1.
Gavezzotti, A. J . Am. Chem. SOC.1985, 107, 962.
Gavezzotti, A. Nouv. J. Chim. 1982, 6,443.
McBridge, J. M. J . Am. Chem. SOC.
1971,93, 6302.
Skinner, K. J.; Blaskiewicz, R. J.; McBridge, J. M. Zsr. J .
Chem. 1972. 10.457.
Jaffe, A. B.; Skinner, K. J.; McBride, J. M. J. Am. Chem. SOC.
1972. 94. 8510.
Karch, N. J.; Koh, E. T.; Whitsel, B. L.; McBride, J. M. J.
Am. Chem. SOC.1975,97, 6729.
McBride, J. M.; Gisler, M. R. Mol. Cryst. Liq. Cryst. 1979,
52. 121.

W.; McBride, J. M. Mol. Cryst. Liq. Cryst. 1979,52,


133.
Walter, D. W.; McBride, J. M. J. Am. Chem. SOC.1981, 103,
7069.
Walter, D. W.; McBride, J. M. J . Am. Chem. SOC.1981, 103,
7064.
McBride, J. M.; Vary, M. W. Tetrahedron 1982, 38, 765.
McBride, J. M. Acc. Chem. Res. 1983, 16, 304.
Craig, D. P.; Sarti-Fantoni, P. Chem. Commun. 1966, 742.
Thomas, J. M.; Williams, J. 0. Chem. Commun. 1967, 432.
Williams, J. 0.;Thomas, J. M. Trans. Faraday SOC.1967,63,
1720.
&$,M.

',

Chemical Reviews, 1987, Vol. 87, No. 2 481

;lotochemical Reactions of Organic Crystals

(283) Cohen, M. D.; Ludmer, Z.; Thomas, J. M.; Williams, J. 0.


Chem. Commun. 1969, 1172.
(284) Cohen, M. D.; Ludmer, Z.; Thomas, J. M.; Williams, J. 0.
Proc. R. SOC.London, A 1971,324,459.
(285) Thomas, J. M. Isr. J. Chem. 1972, 10, 573.
(286) Desvergne, J. P.; Thomas, J. M.; Williams, J. 0.; BouasLaurent. H. J. Chem. So?.. Perkin Trans. 2 1974. 362.
(287) Desver,,,
J. P.; Bouas-Lakent, H.; Lapouyade, R.; Gaultier,
J.; Hauw, C.; Dupuy, F. Mol. Cryst. Liq. Cryst. 1972,19,63.
(288) Thomas, J. M.; Williams, J. 0.;Desvergne, J. P.; Guarini, G.;
Bouas-Laurent, H. J . Chem. SOC.,Perkin Trans. 2 1975,84.
(289) Ramdas, S.; Thomas, J. M.; Goringe, M. J. J. Chem. SOC.
Faraday Trans. 2 1977, 73, 551.
(290) Parkinson, G. M.; Goringe, M. J.; Ramdas, S.; Williams, J. 0.;
Thomas, J. M. J. Chem. Soc., Chem. Commun. 1978, 134.
(291) Ramdas, S.; Jones, W.; Thomas, J. M.; Desvergne, J. P.
Chem. Phys. Lett. 1978, 57, 468.
(292) Tanner, B. K. X-ray Diffraction Topography; Pergamon
Press: Oxford. 1976.
(293) Bart, J. C. J.; Schmidt, G. M. J. Isr. J. Chem. 1971, 9, 429.
(294) Heller, E.; Schmidt, G. M. J. Isr. J. Chem. 1971, 9, 449.
(295) Luther, R.; Weigert, F. 2. Phys. Chem. 1905,51, 297.
(296) Stevens, B.; Dickinson, T.; Sharpe, R. R. Nature (London)
1964, 204, 876.

(297) Chandross, E. A.; Ferguson, J. J. Chem. Phys. 1966,45,3564.


(298) O'Donnel, M. Nature (London) 1968,218,460.
(299) Bouas-Laurent, H.; Lapouyade, R.; Faugere, J. G. C. R. C. R.
Hebd. Seances Acad. Sci., Ser. C 1967, 265, 506.
(300) Ludmer, Z. Chem. Phys. 1977,26, 113.
(301) Ludmer. Z. J. Lumin. 1978. 17. 1.
Ebeid, E. Z. M.; Bridge, N. J. J.'Chem. Soc., Faraday Trans.
1 1984,80, 1131.
Cohen, M. D.; Ron, I.; Schmidt, G. M. J.; Thomas, J. M.
Nature (London) 1969, 224, 167.
Desvergne, J. P.; Chekpo, F.; Bouas-Laurent, H. J . Chem.
SOC..Perkin Trans. 2 1978. 84.
(305) Jones, W.; Thomas, J. M. Prog. Solid State Chem. 1979,12,
101.

(306) Wi*
liams, J. 0. Sci. Prog. Oxford, 1977, 64, 247.
(307) Williams, J. 0.;Thomas, J. M. Surface and Defects Properties of Solids. SDecialist Periodical ReDorts:
. Chemical Socioty: London; 1973; Vol. 2, p 229.
(308) Thomas, J. M.; Williams, J. 0. Surface and Dejects Properties of Solids; Specialist Periodical Reports; Chemical Society: London, 1972; Vol. 1, p 129.
(309) Thomas, J. M.; Williams, J. 0. Prog. Solid State Chem. 1971,
6, 119.

Potrebbero piacerti anche